Handbook of Data Compression, 5th Edition David Salomon Previous editions published under the title “Data Compression: The Complete Reference” Prof. David Salomon (emeritus) Dr. Giovanni Motta Computer Science Dept. Personal Systems Group, Mobility Solutions California State University, Northridge Hewlett-Packard Corp. Northridge, CA 91330-8281 10955 Tantau Ave. USA Cupertino, Califormia 95014-0770 dsalomon@csun.edu gim@ieee.org ISBN 978-1-84882-902-2 e-ISBN 978-1-84882-903-9 DOI 10.1007/10.1007/978-1-84882-903-9 Springer London Dordrecht Heidelberg New York British Library Cataloguing in Publication Data A catalogue record for this book is available from the British Library Library of Congress Control Number: 2009936315 c Springer-Verlag London Limited 2010 Apart from any fair dealing for the purposes of research or private study, or criticism or review, as permitted under the Copyright, Designs and Patents Act 1988, this publication may only be reproduced, stored or transmitted, in any form or by any means, with the prior permission in writing of the publishers, or in the case of reprographic reproduction in accordance with the terms of licenses issued by the Copyright Licensing Agency. Enquiries concerning reproduction outside those terms should be sent to the publishers. The use of registered names, trademarks, etc., in this publication does not imply, even in the absence of a specific statement, that such names are exempt from the relevant laws and regulations and therefore free for general use. The publisher makes no representation, express or implied, with regard to the accuracy of the information contained in this book and cannot accept any legal responsibility or liability for any errors or omissions that may be made. Cover design: eStudio Calamar S.L. Printed on acid-free paper Springer is part of Springer Science+Business Media (www.springer.com) To users of data compression everywhere I love being a writer. What I can’t stand is the paperwork. —Peter De Vries Preface to the New Handbook Gentle Reader. The thick, heavy volume you are holding in your hands was intended to be the fifth edition of Data Compression: The Complete Reference. Instead, its title indicates that this is a handbook of data compression. What makes a book a handbook? What is the difference between a textbook and a handbook? It turns out that “handbook” is one of the many terms that elude precise definition. The many definitions found in dictionaries and reference books vary widely and do more to confuse than to illuminate the reader. Here are a few examples: A concise reference book providing specific information about a subject or location (but this book is not concise). A type of reference work that is intended to provide ready reference (but every reference work should provide ready reference). A pocket reference is intended to be carried at all times (but this book requires big pockets as well as deep ones). A small reference book; a manual (definitely does not apply to this book). General information source which provides quick reference for a given subject area. Handbooks are generally subject-specific (true for this book). Confusing; but we will use the last of these definitions. The aim of this book is to provide a quick reference for the subject of data compression. Judging by the size of the book, the “reference” is certainly there, but what about “quick?” We believe that the following features make this book a quick reference: The detailed index which constitutes 3% of the book. The glossary. Most of the terms, concepts, and techniques discussed throughout the book appear also, albeit briefly, in the glossary. viii Preface The particular organization of the book. Data is compressed by removing redundancies in its original representation, and these redundancies depend on the type of data. Text, images, video, and audio all have different types of redundancies and are best compressed by different algorithms which in turn are based on different approaches. Thus, the book is organized by different data types, with individual chapters devoted to image, video, and audio compression techniques. Some approaches to compression, however, are general and work well on many different types of data, which is why the book also has chapters on variable-length codes, statistical methods, dictionary-based methods, and wavelet methods. The main body of this volume contains 11 chapters and one appendix, all organized in the following categories, basic methods of compression, variable-length codes, statistical methods, dictionary-based methods, methods for image compression, wavelet methods, video compression, audio compression, and other methods that do not conveniently fit into any of the above categories. The appendix discusses concepts of information theory, the theory that provides the foundation of the entire field of data compression. In addition to its use as a quick reference, this book can be used as a starting point to learn more about approaches to and techniques of data compression as well as specific algorithms and their implementations and applications. The broad coverage makes the book as complete as practically possible. The extensive bibliography will be very helpful to those looking for more information on a specific topic. The liberal use of illustrations and tables of data helps to clarify the text. This book is aimed at readers who have general knowledge of computer applications, binary data, and files and want to understand how different types of data can be compressed. The book is not for dummies, nor is it a guide to implementors. Someone who wants to implement a compression algorithm A should have coding experience and should rely on the original publication by the creator of A. In spite of the growing popularity of Internet searching, which often locates quantities of information of questionable quality, we feel that there is still a need for a concise, reliable reference source spanning the full range of the important field of data compression. New to the Handbook The following is a list of the new material in this book (material not included in past editions of Data Compression: The Complete Reference). The topic of compression benchmarks has been added to the Introduction. The paragraphs titled “How to Hide Data” in the Introduction show how data compression can be utilized to quickly and efficiently hide data in plain sight in our computers. Several paragraphs on compression curiosities have also been added to the Introduction. The new Section 1.1.2 shows why irreversible compression may be useful in certain situations. Chapters 2 through 4 discuss the all-important topic of variable-length codes. These chapters discuss basic, advanced, and robust variable-length codes. Many types of VL Preface ix codes are known, they are used by many compression algorithms, have different properties, and are based on different principles. The most-important types of VL codes are prefix codes and codes that include their own length. Section 2.9 on phased-in codes was wrong and has been completely rewritten. An example of the start-step-stop code (2, 2,.) has been added to Section 3.2. Section 3.5 is a description of two interesting variable-length codes dubbed recursive bottom-up coding (RBUC) and binary adaptive sequential coding (BASC). These codes represent compromises between the standard binary (ί) code and the Elias gamma codes. Section 3.28 discusses the original method of interpolative coding whereby dynamic variable-length codes are assigned to a strictly monotonically increasing sequence of integers. Section 5.8 is devoted to the compression of PK (packed) fonts. These are older bitmaps fonts that were developed as part of the huge TEX project. The compression algorithm is not especially efficient, but it provides a rare example of run-length encoding (RLE) without the use of Huffman codes. Section 5.13 is about the Hutter prize for text compression. PAQ (Section 5.15) is an open-source, high-performance compression algorithm and free software that features sophisticated prediction combined with adaptive arithmetic encoding. This free algorithm is especially interesting because of the great interest it has generated and because of the many versions, subversions, and derivatives that have been spun off it. Section 6.3.2 discusses LZR, a variant of the basic LZ77 method, where the lengths of both the search and look-ahead buffers are unbounded. Section 6.4.1 is a description of LZB, an extension of LZSS. It is the result of evaluating and comparing several data structures and variable-length codes with an eye to improving the performance of LZSS. SLH, the topic of Section 6.4.2, is another variant of LZSS. It is a two-pass algorithm where the first pass employs a hash table to locate the best match and to count frequencies, and the second pass encodes the offsets and the raw symbols with Huffman codes prepared from the frequencies counted by the first pass. Most LZ algorithms were developed during the 1980s, but LZPP, the topic of Section 6.5, is an exception. LZPP is a modern, sophisticated algorithm that extends LZSS in several directions and has been inspired by research done and experience gained by many workers in the 1990s. LZPP identifies several sources of redundancy in the various quantities generated and manipulated by LZSS and exploits these sources to obtain better overall compression. Section 6.14.1 is devoted to LZT, an extension of UNIX compress/LZC. The major innovation of LZT is the way it handles a full dictionary. x Preface LZJ (Section 6.17) is an interesting LZ variant. It stores in its dictionary, which can be viewed either as a multiway tree or as a forest, every phrase found in the input. If a phrase is found n times in the input, only one copy is stored in the dictionary. Such behavior tends to fill the dictionary up very quickly, so LZJ limits the length of phrases to a preset parameter h. The interesting, original concept of antidictionary is the topic of Section 6.31. A dictionary-based encoder maintains a list of bits and pieces of the data and employs this list to compress the data. An antidictionary method, on the other hand, maintains a list of strings that do not appear in the data. This generates negative knowledge that allows the encoder to predict with certainty the values of many bits and thus to drop those bits from the output, thereby achieving compression. The important term “pixel” is discussed in Section 7.1, where the reader will discover that a pixel is not a small square, as is commonly assumed, but a mathematical point. Section 7.10.8 discusses the new HD photo (also known as JPEG XR) compression method for continuous-tone still images. ALPC (adaptive linear prediction and classification), is a lossless image compression algorithm described in Section 7.12. ALPC is based on a linear predictor whose coefficients are computed for each pixel individually in a way that can be mimiced by the decoder. Grayscale Two-Dimensional Lempel-Ziv Encoding (GS-2D-LZ, Section 7.18) is an innovative dictionary-based method for the lossless compression of grayscale images. Section 7.19 has been partially rewritten. Section 7.40 is devoted to spatial prediction, a combination of JPEG and fractalbased image compression. A short historical overview of video compression is provided in Section 9.4. The all-important H.264/AVC video compression standard has been extended to allow for a compressed stream that supports temporal, spatial, and quality scalable video coding, while retaining a base layer that is still backward compatible with the original H.264/AVC. This extension is the topic of Section 9.10. The complex and promising VC-1 video codec is the topic of the new, long Section 9.11. The new Section 11.6.4 treats the topic of syllable-based compression, an approach to compression where the basic data symbols are syllables, a syntactic form between characters and words. The commercial compression software known as stuffit has been around since 1987. The methods and algorithms it employs are proprietary, but some information exists in various patents. The new Section 11.16 is an attempt to describe what is publicly known about this software and how it works. There is now a short appendix that presents and explains the basic concepts and terms of information theory. Preface xi We would like to acknowledge the help, encouragement, and cooperation provided by Yuriy Reznik, Matt Mahoney, Mahmoud El-Sakka, Pawel Pylak, Darryl Lovato, Raymond Lau, Cosmin TrutΈa, Derong Bao, and Honggang Qi. They sent information, reviewed certain sections, made useful comments and suggestions, and corrected numerous errors. A special mention goes to David Bryant who wrote Section 10.11. Springer Verlag has created the Springer Handbook series on important scientific and technical subjects, and there can be no doubt that data compression should be included in this category. We are therefore indebted to our editor, Wayne Wheeler, for proposing this project and providing the encouragement and motivation to see it through. The book’s Web site is located at www.DavidSalomon.name. Our email addresses are dsalomon@csun.edu and gim@ieee.org and readers are encouraged to message us with questions, comments, and error corrections. Those interested in data compression in general should consult the short section titled “Joining the Data Compression Community,” at the end of the book, as well as the following resources: http://compression.ca/, http://www-isl.stanford.edu/~gray/iii.html, http://www.hn.is.uec.ac.jp/~arimura/compression_links.html, and http://datacompression.info/. (URLs are notoriously short lived, so search the Internet.) David Salomon Giovanni Motta The preface is usually that part of a book which can most safely be omitted. —William Joyce, Twilight Over England (1940) Preface to the Fourth Edition (This is the Preface to the 4th edition of Data Compression: The Complete Reference, the predecessor of this volume.) I was pleasantly surprised when in November 2005 a message arrived from Wayne Wheeler, the new computer science editor of Springer Verlag, notifying me that he intends to qualify this book as a Springer major reference work (MRW), thereby releasing past restrictions on page counts, freeing me from the constraint of having to compress my style, and making it possible to include important and interesting data compression methods that were either ignored or mentioned in passing in previous editions. These fascicles will represent my best attempt to write a comprehensive account, but computer science has grown to the point where I cannot hope to be an authority on all the material covered in these books. Therefore I’ll need feedback from readers in order to prepare the official volumes later. I try to learn certain areas of computer science exhaustively; then I try to digest that knowledge into a form that is accessible to people who don’t have time for such study. —Donald E. Knuth, http://www-cs-faculty.stanford.edu/~knuth/ (2006) Naturally, all the errors discovered by me and by readers in the third edition have been corrected. Many thanks to all those who bothered to send error corrections, questions, and comments. I also went over the entire book and made numerous additions, corrections, and improvements. In addition, the following new topics have been included in this edition: Tunstall codes (Section 2.6). The advantage of variable-size codes is well known to readers of this book, but these codes also have a downside; they are difficult to work with. The encoder has to accumulate and append several such codes in a short buffer, wait until n bytes of the buffer are full of code bits (where n must be at least 1), write the n bytes on the output, shift the buffer n bytes, and keep track of the location of the last bit placed in the buffer. The decoder has to go through the reverse process. xiv Preface The idea of Tunstall codes is to construct a set of fixed-size codes, each encoding a variable-size string of input symbols. As an aside, the “pod” code (Table 10.29) is also a new addition. Recursive range reduction (3R) (Section 1.7) is a simple coding algorithm due to Yann Guidon that offers decent compression, is easy to program, and its performance is independent of the amount of data to be compressed. LZARI, by Haruhiko Okumura (Section 6.4.3), is an improvement of LZSS. RAR (Section 6.22). The popular RAR software is the creation of Eugene Roshal. RAR has two compression modes, general and special. The general mode employs an LZSS-based algorithm similar to ZIP Deflate. The size of the sliding dictionary in RAR can be varied from 64 Kb to 4 Mb (with a 4 Mb default value) and the minimum match length is 2. Literals, offsets, and match lengths are compressed further by a Huffman coder. An important feature of RAR is an error-control code that increases the reliability of RAR archives while being transmitted or stored. 7-z and LZMA (Section 6.26). LZMA is the main (as well as the default) algorithm used in the popular 7z (or 7-Zip) compression software [7z 06]. Both 7z and LZMA are the creations of Igor Pavlov. The software runs on Windows and is free. Both LZMA and 7z were designed to provide high compression, fast decompression, and low memory requirements for decompression. Stephan Wolf made a contribution to Section 7.34.4. H.264 (Section 9.9). H.264 is an advanced video codec developed by the ISO and the ITU as a replacement for the existing video compression standards H.261, H.262, and H.263. H.264 has the main components of its predecessors, but they have been extended and improved. The only new component in H.264 is a (wavelet based) filter, developed specifically to reduce artifacts caused by the fact that individual macroblocks are compressed separately. Section 10.4 is devoted to the WAVE audio format. WAVE (or simply Wave) is the native file format employed by the Windows opearting system for storing digital audio data. FLAC (Section 10.10). FLAC (free lossless audio compression) is the brainchild of Josh Coalson who developed it in 1999 based on ideas from Shorten. FLAC was especially designed for audio compression, and it also supports streaming and archival of audio data. Coalson started the FLAC project on the well-known sourceforge Web site [sourceforge.flac 06] by releasing his reference implementation. Since then many developers have contributed to improving the reference implementation and writing alternative implementations. The FLAC project, administered and coordinated by Josh Coalson, maintains the software and provides a reference codec and input plugins for several popular audio players. WavPack (Section 10.11, written by David Bryant). WavPack [WavPack 06] is a completely open, multiplatform audio compression algorithm and software that supports three compression modes, lossless, high-quality lossy, and a unique hybrid compression Preface xv mode. It handles integer audio samples up to 32 bits wide and also 32-bit IEEE floatingpoint data [IEEE754 85]. The input stream is partitioned by WavPack into blocks that can be either mono or stereo and are generally 0.5 seconds long (but the length is actually flexible). Blocks may be combined in sequence by the encoder to handle multichannel audio streams. All audio sampling rates are supported by WavPack in all its modes. Monkey’s audio (Section 10.12). Monkey’s audio is a fast, efficient, free, lossless audio compression algorithm and implementation that offers error detection, tagging, and external support. MPEG-4 ALS (Section 10.13). MPEG-4 Audio Lossless Coding (ALS) is the latest addition to the family of MPEG-4 audio codecs. ALS can input floating-point audio samples and is based on a combination of linear prediction (both short-term and longterm), multichannel coding, and efficient encoding of audio residues by means of Rice codes and block codes (the latter are also known as block Gilbert-Moore codes, or BGMC [Gilbert and Moore 59] and [Reznik 04]). Because of this organization, ALS is not restricted to the encoding of audio signals and can efficiently and losslessly compress other types of fixed-size, correlated signals, such as medical (ECG and EEG) and seismic data. AAC (Section 10.15). AAC (advanced audio coding) is an extension of the three layers of MPEG-1 and MPEG-2, which is why it is often called mp4. It started as part of the MPEG-2 project and was later augmented and extended as part of MPEG-4. Apple Computer has adopted AAC in 2003 for use in its well-known iPod, which is why many believe (wrongly) that the acronym AAC stands for apple audio coder. Dolby AC-3 (Section 10.16). AC-3, also known as Dolby Digital, stands for Dolby’s third-generation audio coder. AC-3 is a perceptual audio codec based on the same principles as the three MPEG-1/2 layers and AAC. The new section included in this edition concentrates on the special features of AC-3 and what distinguishes it from other perceptual codecs. Portable Document Format (PDF, Section 11.13). PDF is a popular standard for creating, editing, and printing documents that are independent of any computing platform. Such a document may include text and images (graphics and photos), and its components are compressed by well-known compression algorithms. Section 11.14 (written by Giovanni Motta) covers a little-known but important aspect of data compression, namely how to compress the differences between two files. Hyperspectral data compression (Section 11.15, partly written by Giovanni Motta) is a relatively new and growing field. Hyperspectral data is a set of data items (called pixels) arranged in rows and columns where each pixel is a vector. A home digital camera focuses visible light on a sensor to create an image. In contrast, a camera mounted on a spy satellite (or a satellite searching for minerals and other resources) collects and measures radiation of many wavelegths. The intensity of each wavelength is converted into a number, and the numbers collected from one point on the ground form a vector that becomes a pixel of the hyperspectral data. Another pleasant change is the great help I received from Giovanni Motta, David Bryant, and Cosmin TrutΈa. Each proposed topics for this edition, went over some of xvi Preface the new material, and came up with constructive criticism. In addition, David wrote Section 10.11 and Giovanni wrote Section 11.14 and part of Section 11.15. I would like to thank the following individuals for information about certain topics and for clearing up certain points. Igor Pavlov for help with 7z and LZMA, Stephan Wolf for his contribution, Matt Ashland for help with Monkey’s audio, Yann Guidon for his help with recursive range reduction (3R), Josh Coalson for help with FLAC, and Eugene Roshal for help with RAR. In the first volume of this biography I expressed my gratitude to those individuals and corporate bodies without whose aid or encouragement it would not have been undertaken at all; and to those others whose help in one way or another advanced its progress. With the completion of this volume my obligations are further extended. I should like to express or repeat my thanks to the following for the help that they have given and the premissions they have granted. Christabel Lady Aberconway; Lord Annan; Dr Igor Anrep; . . . —Quentin Bell, Virginia Woolf: A Biography (1972) Currently, the book’s Web site is part of the author’s Web site, which is located at http://www.ecs.csun.edu/~dsalomon/. Domain DavidSalomon.name has been reserved and will always point to any future location of the Web site. The author’s email address is dsalomon@csun.edu, but email sent to anyname@DavidSalomon.name will be forwarded to the author. Those interested in data compression in general should consult the short section titled “Joining the Data Compression Community,” at the end of the book, as well as the following resources: http://compression.ca/, http://www-isl.stanford.edu/~gray/iii.html, http://www.hn.is.uec.ac.jp/~arimura/compression_links.html, and http://datacompression.info/. (URLs are notoriously short lived, so search the Internet). People err who think my art comes easily to me. —Wolfgang Amadeus Mozart Lakeside, California David Salomon Contents Preface to the New Handbook vii Preface to the Fourth Edition xiii Introduction 1 1 Basic Techniques 25 1.1 Intuitive Compression 25 1.2 Run-Length Encoding 31 1.3 RLE Text Compression 31 1.4 RLE Image Compression 36 1.5 Move-to-Front Coding 45 1.6 Scalar Quantization 49 1.7 Recursive Range Reduction 51 2 Basic VL Codes 55 2.1 Codes, Fixed- and Variable-Length 60 2.2 Prefix Codes 62 2.3 VLCs, Entropy, and Redundancy 63 2.4 Universal Codes 68 2.5 The Kraft–McMillan Inequality 69 2.6 Tunstall Code 72 2.7 Schalkwijk’s Coding 74 2.8 Tjalkens–Willems V-to-B Coding 79 2.9 Phased-In Codes 81 2.10 Redundancy Feedback (RF) Coding 85 2.11 Recursive Phased-In Codes 89 2.12 Self-Delimiting Codes 92 xviii Contents 3 Advanced VL Codes 95 3.1 VLCs for Integers 95 3.2 Start-Step-Stop Codes 97 3.3 Start/Stop Codes 99 3.4 Elias Codes 101 3.5 RBUC, Recursive Bottom-Up Coding 107 3.6 Levenstein Code 110 3.7 Even–Rodeh Code 111 3.8 Punctured Elias Codes 112 3.9 Other Prefix Codes 113 3.10 Ternary Comma Code 116 3.11 Location Based Encoding (LBE) 117 3.12 Stout Codes 119 3.13 Boldi–Vigna (.) Codes 122 3.14 Yamamoto’s Recursive Code 125 3.15 VLCs and Search Trees 128 3.16 Taboo Codes 131 3.17 Wang’s Flag Code 135 3.18 Yamamoto Flag Code 137 3.19 Number Bases 141 3.20 Fibonacci Code 143 3.21 Generalized Fibonacci Codes 147 3.22 Goldbach Codes 151 3.23 Additive Codes 157 3.24 Golomb Code 160 3.25 Rice Codes 166 3.26 Subexponential Code 170 3.27 Codes Ending with “1” 171 3.28 Interpolative Coding 172 4 Robust VL Codes 177 4.1 Codes For Error Control 177 4.2 The Free Distance 183 4.3 Synchronous Prefix Codes 184 4.4 Resynchronizing Huffman Codes 190 4.5 Bidirectional Codes 193 4.6 Symmetric Codes 202 4.7 VLEC Codes 204 Contents xix 5 Statistical Methods 211 5.1 Shannon-Fano Coding 211 5.2 Huffman Coding 214 5.3 Adaptive Huffman Coding 234 5.4 MNP5 240 5.5 MNP7 245 5.6 Reliability 247 5.7 Facsimile Compression 248 5.8 PK Font Compression 258 5.9 Arithmetic Coding 264 5.10 Adaptive Arithmetic Coding 276 5.11 The QM Coder 280 5.12 Text Compression 290 5.13 The Hutter Prize 290 5.14 PPM 292 5.15 PAQ 314 5.16 Context-Tree Weighting 320 6 Dictionary Methods 329 6.1 String Compression 331 6.2 Simple Dictionary Compression 333 6.3 LZ77 (Sliding Window) 334 6.4 LZSS 339 6.5 LZPP 344 6.6 Repetition Times 348 6.7 QIC-122 350 6.8 LZX 352 6.9 LZ78 354 6.10 LZFG 358 6.11 LZRW1 361 6.12 LZRW4 364 6.13 LZW 365 6.14 UNIX Compression (LZC) 375 6.15 LZMW 377 6.16 LZAP 378 6.17 LZJ 380 6.18 LZY 383 6.19 LZP 384 6.20 Repetition Finder 391 6.21 GIF Images 394 6.22 RAR and WinRAR 395 6.23 The V.42bis Protocol 398 6.24 Various LZ Applications 399 6.25 Deflate: Zip and Gzip 399 6.26 LZMA and 7-Zip 411 6.27 PNG 416 6.28 XML Compression: XMill 421 xx Contents 6.29 EXE Compressors 423 6.30 Off-Line Dictionary-Based Compression 424 6.31 DCA, Compression with Antidictionaries 430 6.32 CRC 434 6.33 Summary 437 6.34 Data Compression Patents 437 6.35 A Unification 439 7 Image Compression 443 7.1 Pixels 444 7.2 Image Types 446 7.3 Introduction 447 7.4 Approaches to Image Compression 453 7.5 Intuitive Methods 466 7.6 Image Transforms 467 7.7 Orthogonal Transforms 472 7.8 The Discrete Cosine Transform 480 7.9 Test Images 517 7.10 JPEG 520 7.11 JPEG-LS 541 7.12 Adaptive Linear Prediction and Classification 547 7.13 Progressive Image Compression 549 7.14 JBIG 557 7.15 JBIG2 567 7.16 Simple Images: EIDAC 577 7.17 Block Matching 579 7.18 Grayscale LZ Image Compression 582 7.19 Vector Quantization 588 7.20 Adaptive Vector Quantization 598 7.21 Block Truncation Coding 603 7.22 Context-Based Methods 609 7.23 FELICS 612 7.24 Progressive FELICS 615 7.25 MLP 619 7.26 Adaptive Golomb 633 7.27 PPPM 635 7.28 CALIC 636 7.29 Differential Lossless Compression 640 7.30 DPCM 641 7.31 Context-Tree Weighting 646 7.32 Block Decomposition 647 7.33 Binary Tree Predictive Coding 652 7.34 Quadtrees 658 7.35 Quadrisection 676 7.36 Space-Filling Curves 683 7.37 Hilbert Scan and VQ 684 7.38 Finite Automata Methods 695 7.39 Iterated Function Systems 711 7.40 Spatial Prediction 725 7.41 Cell Encoding 729 Contents xxi 8 Wavelet Methods 731 8.1 Fourier Transform 732 8.2 The Frequency Domain 734 8.3 The Uncertainty Principle 737 8.4 Fourier Image Compression 740 8.5 The CWT and Its Inverse 743 8.6 The Haar Transform 749 8.7 Filter Banks 767 8.8 The DWT 777 8.9 Multiresolution Decomposition 790 8.10 Various Image Decompositions 791 8.11 The Lifting Scheme 798 8.12 The IWT 809 8.13 The Laplacian Pyramid 811 8.14 SPIHT 815 8.15 CREW 827 8.16 EZW 827 8.17 DjVu 831 8.18 WSQ, Fingerprint Compression 834 8.19 JPEG 2000 840 9 Video Compression 855 9.1 Analog Video 855 9.2 Composite and Components Video 861 9.3 Digital Video 863 9.4 History of Video Compression 867 9.5 Video Compression 869 9.6 MPEG 880 9.7 MPEG-4 902 9.8 H.261 907 9.9 H.264 910 9.10 H.264/AVC Scalable Video Coding 922 9.11 VC-1 927 10 Audio Compression 953 10.1 Sound 954 10.2 Digital Audio 958 10.3 The Human Auditory System 961 10.4 WAVE Audio Format 969 10.5 ΅-Law and A-Law Companding 971 10.6 ADPCM Audio Compression 977 10.7 MLP Audio 979 10.8 Speech Compression 984 10.9 Shorten 992 10.10 FLAC 996 10.11 WavPack 1007 10.12 Monkey’s Audio 1017 10.13 MPEG-4 Audio Lossless Coding (ALS) 1018 10.14 MPEG-1/2 Audio Layers 1030 10.15 Advanced Audio Coding (AAC) 1055 10.16 Dolby AC-3 1082 xxii Contents 11 Other Methods 1087 11.1 The Burrows-Wheeler Method 1089 11.2 Symbol Ranking 1094 11.3 ACB 1098 11.4 Sort-Based Context Similarity 1105 11.5 Sparse Strings 1110 11.6 Word-Based Text Compression 1121 11.7 Textual Image Compression 1128 11.8 Dynamic Markov Coding 1134 11.9 FHM Curve Compression 1142 11.10 Sequitur 1145 11.11 Triangle Mesh Compression: Edgebreaker 1150 11.12 SCSU: Unicode Compression 1161 11.13 Portable Document Format (PDF) 1167 11.14 File Differencing 1169 11.15 Hyperspectral Data Compression 1180 11.16 Stuffit 1191 A Information Theory 1199 A.1 Information Theory Concepts 1199 Answers to Exercises 1207 Bibliography 1271 Glossary 1303 Joining the Data Compression Community 1329 Index 1331 Content comes first. . . yet excellent design can catch people’s eyes and impress the contents on their memory. —Hideki Nakajima Introduction Giambattista della Porta, a Renaissance scientist sometimes known as the professor of secrets, was the author in 1558 of Magia Naturalis (Natural Magic), a book in which he discusses many subjects, including demonology, magnetism, and the camera obscura [della Porta 58]. The book became tremendously popular in the 16th century and went into more than 50 editions, in several languages beside Latin. The book mentions an imaginary device that has since become known as the “sympathetic telegraph.” This device was to have consisted of two circular boxes, similar to compasses, each with a magnetic needle. Each box was to be labeled with the 26 letters, instead of the usual directions, and the main point was that the two needles were supposed to be magnetized by the same lodestone. Porta assumed that this would somehow coordinate the needles such that when a letter was dialed in one box, the needle in the other box would swing to point to the same letter. Needless to say, such a device does not work (this, after all, was about 300 years before Samuel Morse), but in 1711 a worried wife wrote to the Spectator, a London periodical, asking for advice on how to bear the long absences of her beloved husband. The adviser, Joseph Addison, offered some practical ideas, then mentioned Porta’s device, adding that a pair of such boxes might enable her and her husband to communicate with each other even when they “were guarded by spies and watches, or separated by castles and adventures.” Mr. Addison then added that, in addition to the 26 letters, the sympathetic telegraph dials should contain, when used by lovers, “several entire words which always have a place in passionate epistles.” The message “I love you,” for example, would, in such a case, require sending just three symbols instead of ten. A woman seldom asks advice before she has bought her wedding clothes. —Joseph Addison This advice is an early example of text compression achieved by using short codes for common messages and longer codes for other messages. Even more importantly, this shows how the concept of data compression comes naturally to people who are interested in communications. We seem to be preprogrammed with the idea of sending as little data as possible in order to save time. 2 Introduction Data compression is the process of converting an input data stream (the source stream or the original raw data) into another data stream (the output, the bitstream, or the compressed stream) that has a smaller size. A stream can be a file, a buffer in memory, or individual bits sent on a communications channel. The decades of the 1980s and 1990s saw an exponential decrease in the cost of digital storage. There seems to be no need to compress data when it can be stored inexpensively in its raw format, yet the same two decades have also experienced rapid progress in the development and applications of data compression techniques and algorithms. The following paragraphs try to explain this apparent paradox. Many like to accumulate data and hate to throw anything away. No matter how big a storage device one has, sooner or later it is going to overflow. Data compression is useful because it delays this inevitability. As storage devices get bigger and cheaper, it becomes possible to create, store, and transmit larger and larger data files. In the old days of computing, most files were text or executable programs and were therefore small. No one tried to create and process other types of data simply because there was no room in the computer. In the 1970s, with the advent of semiconductor memories and floppy disks, still images, which require bigger files, became popular. These were followed by audio and video files, which require even bigger files. We hate to wait for data transfers. When sitting at the computer, waiting for a Web page to come in or for a file to download, we naturally feel that anything longer than a few seconds is a long time to wait. Compressing data before it is transmitted is therefore a natural solution. CPU speeds and storage capacities have increased dramatically in the last two decades, but the speed of mechanical components (and therefore the speed of disk input/ output) has increased by a much smaller factor. Thus, it makes sense to store data in compressed form, even if plenty of storage space is still available on a disk drive. Compare the following scenarios: (1) A large program resides on a disk. It is read into memory and is executed. (2) The same program is stored on the disk in compressed form. It is read into memory, decompressed, and executed. It may come as a surprise to learn that the latter case is faster in spite of the extra CPU work involved in decompressing the program. This is because of the huge disparity between the speeds of the CPU and the mechanical components of the disk drive. A similar situation exists with regard to digital communications. Speeds of communications channels, both wired and wireless, are increasing steadily but not dramatically. It therefore makes sense to compress data sent on telephone lines between fax machines, data sent between cellular telephones, and data (such as web pages and television signals) sent to and from satellites. The field of data compression is often called source coding. We imagine that the input symbols (such as bits, ASCII codes, bytes, audio samples, or pixel values) are emitted by a certain information source and have to be coded before being sent to their destination. The source can be memoryless, or it can have memory. In the former case, each symbol is independent of its predecessors. In the latter case, each symbol depends Introduction 3 on some of its predecessors and, perhaps, also on its successors, so they are correlated. A memoryless source is also termed “independent and identically distributed” or IIID. Data compression has come of age in the last 20 years. Both the quantity and the quality of the body of literature in this field provide ample proof of this. However, the need for compressing data has been felt in the past, even before the advent of computers, as the following quotation suggests: I have made this letter longer than usual because I lack the time to make it shorter. —Blaise Pascal There are many known methods for data compression. They are based on different ideas, are suitable for different types of data, and produce different results, but they are all based on the same principle, namely they compress data by removing redundancy from the original data in the source file. Any nonrandom data has some structure, and this structure can be exploited to achieve a smaller representation of the data, a representation where no structure is discernible. The terms redundancy and structure are used in the professional literature, as well as smoothness, coherence, and correlation; they all refer to the same thing. Thus, redundancy is a key concept in any discussion of data compression.  Exercise Intro.1: (Fun) Find English words that contain all five vowels “aeiou” in their original order. In typical English text, for example, the letter E appears very often, while Z is rare (Tables Intro.1 and Intro.2). This is called alphabetic redundancy, and it suggests assigning variable-length codes to the letters, with E getting the shortest code and Z getting the longest code. Another type of redundancy, contextual redundancy, is illustrated by the fact that the letter Q is almost always followed by the letter U (i.e., that in plain English certain digrams and trigrams are more common than others). Redundancy in images is illustrated by the fact that in a nonrandom image, adjacent pixels tend to have similar colors. Section A.1 discusses the theory of information and presents a rigorous definition of redundancy. However, even without a precise definition for this term, it is intuitively clear that a variable-length code has less redundancy than a fixed-length code (or no redundancy at all). Fixed-length codes make it easier to work with text, so they are useful, but they are redundant. The idea of compression by reducing redundancy suggests the general law of data compression, which is to “assign short codes to common events (symbols or phrases) and long codes to rare events.” There are many ways to implement this law, and an analysis of any compression method shows that, deep inside, it works by obeying the general law. Compressing data is done by changing its representation from inefficient (i.e., long) to efficient (short). Compression is therefore possible only because data is normally represented in the computer in a format that is longer than absolutely necessary. The reason that inefficient (long) data representations are used all the time is that they make it easier to process the data, and data processing is more common and more important than data compression. The ASCII code for characters is a good example of a data 4 Introduction Letter Freq. Prob. Letter Freq. Prob. A 51060 0.0721 E 86744 0.1224 B 17023 0.0240 T 64364 0.0908 C 27937 0.0394 I 55187 0.0779 D 26336 0.0372 S 51576 0.0728 E 86744 0.1224 A 51060 0.0721 F 19302 0.0272 O 48277 0.0681 G 12640 0.0178 N 45212 0.0638 H 31853 0.0449 R 45204 0.0638 I 55187 0.0779 H 31853 0.0449 J 923 0.0013 L 30201 0.0426 K 3812 0.0054 C 27937 0.0394 L 30201 0.0426 D 26336 0.0372 M 20002 0.0282 P 20572 0.0290 N 45212 0.0638 M 20002 0.0282 O 48277 0.0681 F 19302 0.0272 P 20572 0.0290 B 17023 0.0240 Q 1611 0.0023 U 16687 0.0235 R 45204 0.0638 G 12640 0.0178 S 51576 0.0728 W 9244 0.0130 T 64364 0.0908 Y 8953 0.0126 U 16687 0.0235 V 6640 0.0094 V 6640 0.0094 X 5465 0.0077 W 9244 0.0130 K 3812 0.0054 X 5465 0.0077 Z 1847 0.0026 Y 8953 0.0126 Q 1611 0.0023 Z 1847 0.0026 J 923 0.0013 Frequencies and probabilities of the 26 letters in a previous edition of this book. The histogram in the background illustrates the byte distribution in the text. Most, but not all, experts agree that the most common letters in English, in order, are ETAOINSHRDLU (normally written as two separate words ETAOIN SHRDLU). However, [Fang 66] presents a different viewpoint. The most common digrams (2-letter combinations) are TH, HE, AN, IN, HA, OR, ND, RE, ER, ET, EA, and OU. The most frequently appearing letters beginning words are S, P, and C, and the most frequent final letters are E, Y, and S. The 11 most common letters in French are ESARTUNILOC. Table Intro.1: Probabilities of English Letters. 0.00 0 50 100 150 200 250 0.05 0.10 0.15 0.20 cr space Relative freq. Byte value uppercase letters and digits lowercase letters Introduction 5 Char. Freq. Prob. Char. Freq. Prob. Char. Freq. Prob. e 85537 0.099293 x 5238 0.006080 F 1192 0.001384 t 60636 0.070387 | 4328 0.005024 H 993 0.001153 i 53012 0.061537 - 4029 0.004677 B 974 0.001131 s 49705 0.057698 ) 3936 0.004569 W 971 0.001127 a 49008 0.056889 ( 3894 0.004520 + 923 0.001071 o 47874 0.055573 T 3728 0.004328 ! 895 0.001039 n 44527 0.051688 k 3637 0.004222 # 856 0.000994 r 44387 0.051525 3 2907 0.003374 D 836 0.000970 h 30860 0.035823 4 2582 0.002997 R 817 0.000948 l 28710 0.033327 5 2501 0.002903 M 805 0.000934 c 26041 0.030229 6 2190 0.002542 ; 761 0.000883 d 25500 0.029601 I 2175 0.002525 / 698 0.000810 m 19197 0.022284 ^ 2143 0.002488 N 685 0.000795 \ 19140 0.022218 : 2132 0.002475 G 566 0.000657 p 19055 0.022119 A 2052 0.002382 j 508 0.000590 f 18110 0.021022 9 1953 0.002267 @ 460 0.000534 u 16463 0.019111 [ 1921 0.002230 Z 417 0.000484 b 16049 0.018630 C 1896 0.002201 J 415 0.000482 . 12864 0.014933 ] 1881 0.002183 O 403 0.000468 1 12335 0.014319 ’ 1876 0.002178 V 261 0.000303 g 12074 0.014016 S 1871 0.002172 X 227 0.000264 0 10866 0.012613 _ 1808 0.002099 U 224 0.000260 , 9919 0.011514 7 1780 0.002066 ? 177 0.000205 & 8969 0.010411 8 1717 0.001993 K 175 0.000203 y 8796 0.010211 ‘ 1577 0.001831 % 160 0.000186 w 8273 0.009603 = 1566 0.001818 Y 157 0.000182 $ 7659 0.008891 P 1517 0.001761 Q 141 0.000164 } 6676 0.007750 L 1491 0.001731 > 137 0.000159 { 6676 0.007750 q 1470 0.001706 * 120 0.000139 v 6379 0.007405 z 1430 0.001660 < 99 0.000115 2 5671 0.006583 E 1207 0.001401 ” 8 0.000009 Frequencies and probabilities of the 93 most-common characters in a prepublication previous edition of this book, containing 861,462 characters. See Figure Intro.3 for the Mathematica code. Table Intro.2: Frequencies and Probabilities of Characters. 6 Introduction representation that is longer than absolutely necessary. It uses 7-bit codes because fixed-size codes are easy to work with. A variable-size code, however, would be more efficient, since certain characters are used more than others and so could be assigned shorter codes. In a world where data is always represented by its shortest possible format, there would therefore be no way to compress data. Instead of writing books on data compression, authors in such a world would write books on how to determine the shortest format for different types of data. fpc = OpenRead["test.txt"]; g = 0; ar = Table[{i, 0}, {i, 256}]; While[0 == 0, g = Read[fpc, Byte]; (* Skip space, newline & backslash *) If[g==10||g==32||g==92, Continue[]]; If[g==EndOfFile, Break[]]; ar[[g, 2]]++] (* increment counter *) Close[fpc]; ar = Sort[ar, #1[[2]] > #2[[2]] &]; tot = Sum[ ar[[i,2]], {i,256}] (* total chars input *) Table[{FromCharacterCode[ar[[i,1]]],ar[[i,2]],ar[[i,2]]/N[tot,4]}, {i,93}] (* char code, freq., percentage *) TableForm[%] Figure Intro.3: Code for Table Intro.2. A Word to the Wise . . . The main aim of the field of data compression is, of course, to develop methods for better and faster compression. However, one of the main dilemmas of the art of data compression is when to stop looking for better compression. Experience shows that fine-tuning an algorithm to squeeze out the last remaining bits of redundancy from the data gives diminishing returns. Modifying an algorithm to improve compression by 1% may increase the run time by 10% and the complexity of the program by more than that. A good way out of this dilemma was taken by Fiala and Greene (Section 6.10). After developing their main algorithms A1 and A2, they modified them to produce less compression at a higher speed, resulting in algorithms B1 and B2. They then modified A1 and A2 again, but in the opposite direction, sacrificing speed to get slightly better compression. The principle of compressing by removing redundancy also answers the following question: Why is it that an already compressed file cannot be compressed further? The Introduction 7 answer, of course, is that such a file has little or no redundancy, so there is nothing to remove. An example of such a file is random text. In such text, each letter occurs with equal probability, so assigning them fixed-size codes does not add any redundancy. When such a file is compressed, there is no redundancy to remove. (Another answer is that if it were possible to compress an already compressed file, then successive compressions would reduce the size of the file until it becomes a single byte, or even a single bit. This, of course, is ridiculous since a single byte cannot contain the information present in an arbitrarily large file.) In spite of the arguments above and the proof below, claims of recursive compression appear from time to time in the Internet. These are either checked and proved wrong or disappear silently. Reference [Barf 08], is a joke intended to amuse (and temporarily confuse) readers. A careful examination of this “claim” shows that any gain achieved by recursive compression of the Barf software is offset (perhaps more than offset) by the long name of the output file generated. The reader should also consult page 1132 for an interesting twist on the topic of compressing random data. Definition of barf (verb): to vomit; purge; cast; sick; chuck; honk; throw up. Since random data has been mentioned, let’s say a few more words about it. Normally, it is rare to have a file with random data, but there is at least one good example— an already compressed file. Someone owning a compressed file normally knows that it is already compressed and would not attempt to compress it further, but there may be exceptions and one of them is data transmission by modems. Modern modems include hardware to automatically compress the data they send, and if that data is already compressed, there will not be further compression. There may even be expansion. This is why a modem should monitor the compression ratio “on the fly,” and if it is low, it should stop compressing and should send the rest of the data uncompressed. The V.42bis protocol (Section 6.23) is a good example of this technique. Before we prove the impossibility of recursive compression, here is an interesting twist on this concept. Several algorithms, such as JPEG, LZW, and MPEG, have long become de facto standards and are commonly used in web sites and in our computers. The field of data compression, however, is rapidly advancing and new, sophisticated methods are continually being developed. Thus, it is possible to take a compressed file, say JPEG, decompress it, and recompress it with a more efficient method. On the outside, this would look like recursive compression and may become a marketing tool for new, commercial compression software. The Stuffit software for the Macintosh platform (Section 11.16) does just that. It promises to compress already-compressed files and in many cases, it does! The following simple argument illustrates the essence of the statement “Data compression is achieved by reducing or removing redundancy in the data.” The argument shows that most data files cannot be compressed, no matter what compression method is used. This seems strange at first because we compress our data files all the time. The point is that most files cannot be compressed because they are random or close to random and therefore have no redundancy. The (relatively) few files that can be compressed are the ones that we want to compress; they are the files we use all the time. They have redundancy, are nonrandom, and are therefore useful and interesting. 8 Introduction Here is the argument. Given two different files A and B that are compressed to files C and D, respectively, it is clear that C and D must be different. If they were identical, there would be no way to decompress them and get back file A or file B. Suppose that a file of size n bits is given and we want to compress it efficiently. Any compression method that can compress this file to, say, 10 bits would be welcome. Even compressing it to 11 bits or 12 bits would be great. We therefore (somewhat arbitrarily) assume that compressing such a file to half its size or better is considered good compression. There are 2n n-bit files and they would have to be compressed into 2n different files of sizes less than or equal to n/2. However, the total number of these files is N = 1+2+4+· · · + 2n/2 = 21+n/2 - 1 . 21+n/2, so only N of the 2n original files have a chance of being compressed efficiently. The problem is that N is much smaller than 2n. Here are two examples of the ratio between these two numbers. For n = 100 (files with just 100 bits), the total number of files is 2100 and the number of files that can be compressed efficiently is 251. The ratio of these numbers is the ridiculously small fraction 2-49 . 1.78Χ10-15. For n = 1000 (files with just 1000 bits, about 125 bytes), the total number of files is 21000 and the number of files that can be compressed efficiently is 2501. The ratio of these numbers is the incredibly small fraction 2-499 . 9.82Χ10-91. Most files of interest are at least some thousands of bytes long. For such files, the percentage of files that can be efficiently compressed is so small that it cannot be computed with floating-point numbers even on a supercomputer (the result comes out as zero). The 50% figure used here is arbitrary, but even increasing it to 90% isn’t going to make a significant difference. Here is why. Assuming that a file of n bits is given and that 0.9n is an integer, the number of files of sizes up to 0.9n is 20 + 21 + · · · + 20.9n = 21+0.9n - 1 . 21+0.9n. For n = 100, there are 2100 files and 21+90 = 291 of them can be compressed well. The ratio of these numbers is 291/2100 = 2-9 . 0.00195. For n = 1000, the corresponding fraction is 2901/21000 = 2-99 . 1.578Χ10-30. These are still extremely small fractions. It is therefore clear that no compression method can hope to compress all files or even a significant percentage of them. In order to compress a data file, the compression algorithm has to examine the data, find redundancies in it, and try to remove them. The redundancies in data depend on the type of data (text, images, audio, etc.), which is why a new compression method has to be developed for each specific type of data and it performs best on this type. There is no such thing as a universal, efficient data compression algorithm. Data compression has become so important that some researchers (see, for example, [Wolff 99]) have proposed the SP theory (for “simplicity” and “power”), which suggests that all computing is compression! Specifically, it says: Data compression may be interpreted as a process of removing unnecessary complexity (redundancy) in information, and thereby maximizing simplicity while preserving as much as possible of its nonredundant descriptive power. SP theory is based on the following conjectures: Introduction 9 All kinds of computing and formal reasoning may usefully be understood as information compression by pattern matching, unification, and search. The process of finding redundancy and removing it may always be understood at a fundamental level as a process of searching for patterns that match each other, and merging or unifying repeated instances of any pattern to make one. This book discusses many compression methods, some suitable for text and others for graphical data (still images or video) or for audio. Most methods are classified into four categories: run length encoding (RLE), statistical methods, dictionary-based (sometimes called LZ) methods, and transforms. Chapters 1 and 11 describe methods based on other principles. Before delving into the details, we discuss important data compression terms. A compressor or encoder is a program that compresses the raw data in the input stream and creates an output stream with compressed (low-redundancy) data. A decompressor or decoder converts in the opposite direction. Note that the term encoding is very general and has several meanings, but since we discuss only data compression, we use the name encoder to mean data compressor. The term codec is often used to describe both the encoder and the decoder. Similarly, the term companding is short for “compressing/expanding.” The term stream is used throughout this book instead of file. Stream is a more general term because the compressed data may be transmitted directly to the decoder, instead of being written to a file and saved. Also, the data to be compressed may be downloaded from a network instead of being input from a file. For the original input stream, we use the terms unencoded, raw, or original data. The contents of the final, compressed, stream are considered the encoded or compressed data. The term bitstream is also used in the literature to indicate the compressed stream. The Gold Bug Here, then, we have, in the very beginning, the groundwork for something more than a mere guess. The general use which may be made of the table is obvious—but, in this particular cipher, we shall only very partially require its aid. As our predominant character is 8, we will commence by assuming it as the “e” of the natural alphabet. To verify the supposition, let us observe if the 8 be seen often in couples—for “e” is doubled with great frequency in English—in such words, for example, as “meet,” “fleet,” “speed,” “seen,” “been,” “agree,” etc. In the present instance we see it doubled no less than five times, although the cryptograph is brief. —Edgar Allan Poe A nonadaptive compression method is rigid and does not modify its operations, its parameters, or its tables in response to the particular data being compressed. Such a method is best used to compress data that is all of a single type. Examples are 10 Introduction the Group 3 and Group 4 methods for facsimile compression (Section 5.7). They are specifically designed for facsimile compression and would do a poor job compressing any other data. In contrast, an adaptive method examines the raw data and modifies its operations and/or its parameters accordingly. An example is the adaptive Huffman method of Section 5.3. Some compression methods use a 2-pass algorithm, where the first pass reads the input stream to collect statistics on the data to be compressed, and the second pass does the actual compressing using parameters determined by the first pass. Such a method may be called semiadaptive. A data compression method can also be locally adaptive, meaning it adapts itself to local conditions in the input stream and varies this adaptation as it moves from area to area in the input. An example is the move-to-front method (Section 1.5). Lossy/lossless compression: Certain compression methods are lossy. They achieve better compression at the price of losing some information. When the compressed stream is decompressed, the result is not identical to the original data stream. Such a method makes sense especially in compressing images, video, or audio. If the loss of data is small, we may not be able to tell the difference. In contrast, text files, especially files containing computer programs, may become worthless if even one bit gets modified. Such files should be compressed only by a lossless compression method. [Two points should be mentioned regarding text files: (1) If a text file contains the source code of a program, consecutive blank spaces can often be replaced by a single space. (2) When the output of a word processor is saved in a text file, the file may contain information about the different fonts used in the text. Such information may be discarded if the user is interested in saving just the text.] Cascaded compression: The difference between lossless and lossy codecs can be illuminated by considering a cascade of compressions. Imagine a data file A that has been compressed by an encoder X, resulting in a compressed file B. It is possible, although pointless, to pass B through another encoder Y , to produce a third compressed file C. The point is that if methods X and Y are lossless, then decoding C by Y will produce an exact B, which when decoded by X will yield the original file A. However, if any of the compression algorithms is lossy, then decoding C by Y may produce a file B different from B. Passing B through X may produce something very different from A and may also result in an error, because X may not be able to read B. Perceptive compression: A lossy encoder must take advantage of the special type of data being compressed. It should delete only data whose absence would not be detected by our senses. Such an encoder must therefore employ algorithms based on our understanding of psychoacoustic and psychovisual perception, so it is often referred to as a perceptive encoder. Such an encoder can be made to operate at a constant compression ratio, where for each x bits of raw data, it outputs y bits of compressed data. This is convenient in cases where the compressed stream has to be transmitted at a constant rate. The trade-off is a variable subjective quality. Parts of the original data that are difficult to compress may, after decompression, look (or sound) bad. Such parts may require more than y bits of output for x bits of input. Symmetrical compression is the case where the compressor and decompressor employ basically the same algorithm but work in “opposite” directions. Such a method Introduction 11 makes sense for general work, where the same number of files are compressed as are decompressed. In an asymmetric compression method, either the compressor or the decompressor may have to work significantly harder. Such methods have their uses and are not necessarily bad. A compression method where the compressor executes a slow, complex algorithm and the decompressor is simple is a natural choice when files are compressed into an archive, where they will be decompressed and used very often. The opposite case is useful in environments where files are updated all the time and backups are made. There is a small chance that a backup file will be used, so the decompressor isn’t used very often. Like the ski resort full of girls hunting for husbands and husbands hunting for girls, the situation is not as symmetrical as it might seem. —Alan Lindsay Mackay, lecture, Birckbeck College, 1964  Exercise Intro.2: Give an example of a compressed file where good compression is important but the speed of both compressor and decompressor isn’t important. Many modern compression methods are asymmetric. Often, the formal description (the standard) of such a method specifies the decoder and the format of the compressed stream, but does not discuss the operation of the encoder. Any encoder that generates a correct compressed stream is considered compliant, as is also any decoder that can read and decode such a stream. The advantage of such a description is that anyone is free to develop and implement new, sophisticated algorithms for the encoder. The implementor need not even publish the details of the encoder and may consider it proprietary. If a compliant encoder is demonstrably better than competing encoders, it may become a commercial success. In such a scheme, the encoder is considered algorithmic, while the decoder, which is normally much simpler, is termed deterministic. A good example of this approach is the MPEG-1 audio compression method (Section 10.14). A data compression method is called universal if the compressor and decompressor do not know the statistics of the input stream. A universal method is optimal if the compressor can produce compression ratios that asymptotically approach the entropy of the input stream for long inputs. The term file differencing refers to any method that locates and compresses the differences between two files. Imagine a file A with two copies that are kept by two users. When a copy is updated by one user, it should be sent to the other user, to keep the two copies identical. Instead of sending a copy of A, which may be big, a much smaller file containing just the differences, in compressed format, can be sent and used at the receiving end to update the copy of A. Section 11.14.2 discusses some of the details and shows why compression can be considered a special case of file differencing. Note that the term differencing is used in Section 1.3.1 to describe an entirely different compression method. Most compression methods operate in the streaming mode, where the codec inputs a byte or several bytes, processes them, and continues until an end-of-file is sensed. Some methods, such as Burrows-Wheeler transform (Section 11.1), work in the block mode, where the input stream is read block by block and each block is encoded separately. The 12 Introduction block size should be a user-controlled parameter, since its size may significantly affect the performance of the method. Most compression methods are physical. They look only at the bits in the input stream and ignore the meaning of the data items in the input (e.g., the data items may be words, pixels, or audio samples). Such a method translates one bitstream into another, shorter bitstream. The only way to make sense of the output stream (to decode it) is by knowing how it was encoded. Some compression methods are logical. They look at individual data items in the source stream and replace common items with short codes. A logical method is normally special purpose and can operate successfully on certain types of data only. The pattern substitution method described on page 35 is an example of a logical compression method. Compression performance: Several measures are commonly used to express the performance of a compression method. 1. The compression ratio is defined as Compression ratio = size of the output stream size of the input stream . A value of 0.6 means that the data occupies 60% of its original size after compression. Values greater than 1 imply an output stream bigger than the input stream (negative compression). The compression ratio can also be called bpb (bit per bit), since it equals the number of bits in the compressed stream needed, on average, to compress one bit in the input stream. In modern, efficient text compression methods, it makes sense to talk about bpc (bits per character)—the number of bits it takes, on average, to compress one character in the input stream. Two more terms should be mentioned in connection with the compression ratio. The term bitrate (or “bit rate”) is a general term for bpb and bpc. Thus, the main goal of data compression is to represent any given data at low bit rates. The term bit budget refers to the functions of the individual bits in the compressed stream. Imagine a compressed stream where 90% of the bits are variable-size codes of certain symbols, and the remaining 10% are used to encode certain tables. The bit budget for the tables is 10%. 2. The inverse of the compression ratio is called the compression factor: Compression factor = size of the input stream size of the output stream. In this case, values greater than 1 indicate compression and values less than 1 imply expansion. This measure seems natural to many people, since the bigger the factor, the better the compression. This measure is distantly related to the sparseness ratio, a performance measure discussed in Section 8.6.2. 3. The expression 100 Χ (1 - compression ratio) is also a reasonable measure of compression performance. A value of 60 means that the output stream occupies 40% of its original size (or that the compression has resulted in savings of 60%). Introduction 13 4. In image compression, the quantity bpp (bits per pixel) is commonly used. It equals the number of bits needed, on average, to compress one pixel of the image. This quantity should always be compared with the bpp before compression. 5. The compression gain is defined as 100 loge reference size compressed size , where the reference size is either the size of the input stream or the size of the compressed stream produced by some standard lossless compression method. For small numbers x, it is true that loge(1 + x) . x, so a small change in a small compression gain is very similar to the same change in the compression ratio. Because of the use of the logarithm, two compression gains can be compared simply by subtracting them. The unit of the compression gain is called percent log ratio and is denoted by .–. . 6. The speed of compression can be measured in cycles per byte (CPB). This is the average number of machine cycles it takes to compress one byte. This measure is important when compression is done by special hardware. 7. Other quantities, such as mean square error (MSE) and peak signal to noise ratio (PSNR), are used to measure the distortion caused by lossy compression of images and movies. Section 7.4.2 provides information on those. 8. Relative compression is used to measure the compression gain in lossless audio compression methods, such as MLP (Section 10.7). This expresses the quality of compression by the number of bits each audio sample is reduced. Name Size Description Type bib 111,261 A bibliography in UNIX refer format Text book1 768,771 Text of T. Hardy’s Far From the Madding Crowd Text book2 610,856 Ian Witten’s Principles of Computer Speech Text geo 102,400 Geological seismic data Data news 377,109 A Usenet news file Text obj1 21,504 VAX object program Obj obj2 246,814 Macintosh object code Obj paper1 53,161 A technical paper in troff format Text paper2 82,199 Same Text pic 513,216 Fax image (a bitmap) Image progc 39,611 A source program in C Source progl 71,646 A source program in LISP Source progp 49,379 A source program in Pascal Source trans 93,695 Document teaching how to use a terminal Text Table Intro.4: The Calgary Corpus. The Calgary Corpus is a set of 18 files traditionally used to test data compression algorithms and implementations. They include text, image, and object files, for a total 14 Introduction of more than 3.2 million bytes (Table Intro.4). The corpus can be downloaded from [Calgary 06]. The Canterbury Corpus (Table Intro.5) is another collection of files introduced in 1997 to provide an alternative to the Calgary corpus for evaluating lossless compression methods. The following concerns led to the new corpus: 1. The Calgary corpus has been used by many researchers to develop, test, and compare many compression methods, and there is a chance that new methods would unintentionally be fine-tuned to that corpus. They may do well on the Calgary corpus documents but poorly on other documents. 2. The Calgary corpus was collected in 1987 and is getting old. “Typical” documents change over a period of decades (e.g., html documents did not exist in 1987), and any body of documents used for evaluation purposes should be examined from time to time. 3. The Calgary corpus is more or less an arbitrary collection of documents, whereas a good corpus for algorithm evaluation should be selected carefully. The Canterbury corpus started with about 800 candidate documents, all in the public domain. They were divided into 11 classes, representing different types of documents. A representative “average” document was selected from each class by compressing every file in the class using different methods and selecting the file whose compression was closest to the average (as determined by statistical regression). The corpus is summarized in Table Intro.5 and can be obtained from [Canterbury 06]. Description File name Size (bytes) English text (Alice in Wonderland) alice29.txt 152,089 Fax images ptt5 513,216 C source code fields.c 11,150 Spreadsheet files kennedy.xls 1,029,744 SPARC executables sum 38,666 Technical document lcet10.txt 426,754 English poetry (“Paradise Lost”) plrabn12.txt 481,861 HTML document cp.html 24,603 LISP source code grammar.lsp 3,721 GNU manual pages xargs.1 4,227 English play (As You Like It) asyoulik.txt 125,179 Complete genome of the E. coli bacterium E.Coli 4,638,690 The King James version of the Bible bible.txt 4,047,392 The CIA World Fact Book world192.txt 2,473,400 Table Intro.5: The Canterbury Corpus. The last three files constitute the beginning of a random collection of larger files. More files are likely to be added to it. The Calgary challenge [Calgary challenge 08], is a contest to compress the Calgary corpus. It was started in 1996 by Leonid Broukhis and initially attracted a number Introduction 15 of contestants. In 2005, Alexander Ratushnyak achieved the current record of 596,314 bytes, using a variant of PAsQDa with a tiny dictionary of about 200 words. Currently (late 2008), this challenge seems to have been preempted by the bigger prizes offered by the Hutter prize, and has featured no activity since 2005. Here is part of the original challenge as it appeared in [Calgary challenge 08]. I, Leonid A. Broukhis, will pay the amount of (759, 881.00 - X)/333 US dollars (but not exceeding $1001, and no less than $10.01 “ten dollars and one cent”) to the first person who sends me an archive of length X bytes, containing an executable and possibly other files, where the said executable file, run repeatedly with arguments being the names of other files contained in the original archive file one at a time (or without arguments if no other files are present) on a computer with no permanent storage or communication devices accessible to the running process(es) produces 14 new files, so that a 1-to-1 relationship of bytewise identity may be established between those new files and the files in the original Calgary corpus. (In other words, “solid” mode, as well as shared dictionaries/models and other tune-ups specific for the Calgary Corpus are allowed.) I will also pay the amount of (777, 777.00 - Y )/333 US dollars (but not exceeding $1001, and no less than $0.01 “zero dollars and one cent”) to the first person who sends me an archive of length Y bytes, containing an executable and exactly 14 files, where the said executable file, run with standard input taken directly (so that the stdin is seekable) from one of the 14 other files and the standard output directed to a new file, writes data to standard output so that the data being output matches one of the files in the original Calgary corpus and a 1-to-1 relationship may be established between the files being given as standard input and the files in the original Calgary corpus that the standard output matches. Moreover, after verifying the above requirements, an arbitrary file of size between 500 KB and 1 MB will be sent to the author of the decompressor to be compressed and sent back. The decompressor must handle that file correctly, and the compression ratio achieved on that file must be not worse that within 10% of the ratio achieved by gzip with default settings. (In other words, the compressor must be, roughly speaking, “general purpose.”) The probability model. This concept is important in statistical data compression methods. In such a method, a model for the data has to be constructed before compression can begin. A typical model may be built by reading the entire input stream, counting the number of times each symbol appears (its frequency of occurrence), and computing the probability of occurrence of each symbol. The data stream is then input again, symbol by symbol, and is compressed using the information in the probability model. A typical model is shown in Table 5.42, page 266. Reading the entire input stream twice is slow, which is why practical compression methods use estimates, or adapt themselves to the data as it is being input and compressed. It is easy to scan large quantities of, say, English text and calculate the frequencies and probabilities of every character. This information can later serve as an approximate model for English text and can be used by text compression methods to compress any English text. It is also possible to start by assigning equal probabilities to 16 Introduction all the symbols in an alphabet, then reading symbols and compressing them, and, while doing that, also counting frequencies and changing the model as compression progresses. This is the principle behind the various adaptive compression methods. Source. A source of data items can be a file stored on a disk, a file that is input from outside the computer, text input from a keyboard, or a program that generates data symbols to be compressed or processed in some way. In a memoryless source, the probability of occurrence of a data symbol does not depend on its context. The term i.i.d. (independent and identically distributed) refers to a set of sources that have the same probability distribution and are mutually independent. Alphabet. This is the set of symbols that an application has to deal with. An alphabet may consist of the 128 ASCII codes, the 256 8-bit bytes, the two bits, or any other set of symbols. Random variable. This is a function that maps the results of random experiments to numbers. For example, selecting many people and measuring their heights is a random variable. The number of occurrences of each height can be used to compute the probability of that height, so we can talk about the probability distribution of the random variable (the set of probabilities of the heights). A special important case is a discrete random variable. The set of all values that such a variable can assume is finite or countably infinite. Compressed stream (or encoded stream). A compressor (or encoder) compresses data and generates a compressed stream. This is often a file that is written on a disk or is stored in memory. Sometimes, however, the compressed stream is a string of bits that are transmitted over a communications line. [End of data compression terms.] The concept of data reliability and integrity (page 247) is in some sense the opposite of data compression. Nevertheless, the two concepts are often related since any good data compression program should generate reliable code and so should be able to use error-detecting and error-correcting codes. Compression benchmarks Research in data compression, as in many other areas of computer science, concentrates on finding new algorithms and improving existing ones. In order to prove its value, however, an algorithm has to be implemented and tested. Thus, every researcher, programmer, and developer compares a new algorithm to older, well-established and known methods, and draws conclusions about its performance. In addition to these tests, workers in the field of compression continually conduct extensive benchmarks, where many algorithms are run on the same set of data files and the results are compared and analyzed. This short section describes a few independent compression benchmarks. Perhaps the most-important fact about these benchmarks is that they generally restrict themselves to compression ratios. Thus, a winner in such a benchmark may not be the best choice for general, everyday use, because it may be slow, may require large memory space, and may be expensive or protected by patents. Benchmarks for Introduction 17 compression speed are rare, because it is difficult to accurately measure the run time of an executable program (i.e., a program whose source code is unavailable). Another drawback of benchmarking is that their data files are generally publicly known, so anyone interested in record breaking for its own sake may tune an existing algorithm to the particular data files used by a benchmark and in this way achieve the smallest (but nevertheless meaningless) compression ratio. From the Dictionary Benchmark: A standard by which something can be measured or judged; “his painting sets the benchmark of quality.” In computing, a benchmark is the act of running a computer program, a set of programs, or other operations, in order to assess the relative performance of an object, normally by running a number of standard tests and trials against it. The term benchmark originates from the chiseled horizontal marks that surveyors made in stone structures, into which an angle-iron could be placed to form a “bench” for a leveling rod, thus ensuring that a leveling rod could be accurately repositioned in the same place in future. The following independent benchmarks compare the performance (compression ratios but generally not speeds, which are difficult to measure) of many algorithms and their implementations. Surprisingly, the results often indicate that the winner comes from the family of context-mixing compression algorithms. Such an algorithm employs several models of the data to predict the next data symbol, and then combines the predictions in some way to end up with a probability for the next symbol. The symbol and its computed probability are then sent to an adaptive arithmetic coder, to be encoded. Included in this family of lossless algorithms are the many versions and derivatives of PAQ (Section 5.15) as well as many other, less well-known methods. The Maximum Compression Benchmark, managed byWerner Bergmans. This suite of tests (described in [Bergmans 08]) was started in 2003 and is still frequently updated. The goal is to discover the best compression ratios for several different types of data, such as text, images, and executable code. Every algorithm included in the tests is first tuned by setting any switches and parameters to the values that yield best performance. The owner of this benchmark prefers command line (console) compression programs over GUI ones. At the time of writing (late 2008), more than 150 programs have been tested on several large collections of test files. Most of the top-ranked algorithms are of the context mixing type. Special mention goes to PAsQDa 4.1b and WinRK 2.0.6/pwcm. The latest update to this benchmark reads as follows: 28-September-2008: Added PAQ8P, 7-Zip 4.60b, FreeARC 0.50a (June 23 2008), Tornado 0.4a, M1 0.1a, BZP 0.3, NanoZIP 0.04a, Blizzard 0.24b and WinRAR 3.80b5 (MFC to do for WinRAR and 7-Zip). PAQ8P manages to squeeze out an additional 12 KB from the BMP file, further increasing the gap to the number 2 in the SFC benchmark; newcomer NanoZIP takes 6th place in SFC!. In the MFC benchmark PAQ8 now takes a huge lead over WinRK 3.0.3, but WinRK 3.1.2 is on the todo list to be tested. To be continued. . . . Johan de Bock started the UCLC (ultimate command-line compressors) benchmark 18 Introduction project [UCLC 08]. A wide variety of tests are performed to compare the latest state-ofthe- art command line compressors. The only feature being compared is the compression ratio; run-time and memory requirements are ignored. More than 100 programs have been tested over the years, with WinRK and various versions of PAQ declared the best performers (except for audio and grayscale images, where the records were achieved by specialized algorithms). The EmilCont benchmark [Emilcont 08] is managed by Berto Destasio. At the time of writing, the latest update of this site dates back to March 2007. EmilCont tests hundreds of algorithms on a confidential set of data files that include text, images, audio, and executable code. As usual, WinRK and PAQ variants are among the record holders, followed by SLIM. An important feature of these benchmarks is that the test data will not be released to avoid the unfortunate practice of compression writers tuning their programs to the benchmarks. —From [Emilcont 08] The Archive Comparison Test (ACT), maintained by Jeff Gilchrist [Gilchrist 08], is a set of benchmarks designed to demonstrate the state of the art in lossless data compression. It contains benchmarks on various types of data for compression speed, decompression speed, and compression ratio. The site lists the results of comparing 162 DOS/Windows programs, eight Macintosh programs, and 11 JPEG compressors. However, the tests are old. Many were performed in or before 2002 (except the JPEG tests, which took place in 2007). In [Ratushnyak 08], Alexander Ratushnyak reports the results of hundreds of speed tests done in 2001. Site [squeezemark 09] is devoted to benchmarks of lossless codecs. It is kept up to date by its owner, Stephan Busch, and has many pictures of compression pioneers. How to Hide Data Here is an unforeseen, unexpected application of data compression. For a long time I have been thinking about how to best hide a sensitive data file in a computer, while still having it ready for use at a short notice. Here is what we came up with. Given a data file A, consider the following steps: 1. Compress A. The result is a file B that is small and also seems random. This has two advantages (1) the remaining steps encrypt and hide small files and (2) the next step encrypts a random file, thereby making it difficult to break the encryption simply by checking every key. 2. Encrypt B with a secret key to obtain file C. A would-be codebreaker may attempt to decrypt C by writing a program that loops and tries every key, but here is the rub. Each time a key is tried, someone (or something) has to check the result. If the result looks meaningful, it may be the decrypted file B, but if the result seems random, the loop should continue. At the end of the loop; frustration. Introduction 19 3. Hide C inside a cover file D to obtain a large file E. Use one of the many steganographic methods for this (notice that many such methods depend on secret keys). One reference for steganography is [Salomon 03], but today there may be better texts. 4. Hide E in plain sight in your computer by changing its name and placing it in a large folder together with hundreds of other, unfamiliar files. A good idea may be to change the file name to msLibPort.dll (or something similar that includes MS and other familiar-looking terms) and place it in one of the many large folders created and used exclusively by Windows or any other operating system. If files in this folder are visible, do not make your file invisible. Anyone looking inside this folder will see hundreds of unfamiliar files and will have no reason to suspect msLibPort.dll. Even if this happens, an opponent would have a hard time guessing the three steps above (unless he has read these paragraphs) and the keys used. If file E is large (perhaps more than a few Gbytes), it should be segmented into several smaller files and each hidden in plain sight as described above. This step is important because there are utilities that identify large files and they may attract unwanted attention to your large E. For those who require even greater privacy, here are a few more ideas. (1) A password can be made strong by including in it special characters such §, Ά, †, and ‡. These can be typed with the help of special modifier keys found on most keyboards. (2) Add a step between steps 1 and 2 where file B is recompressed by any compression method. This will not decrease the size of B but will defeat anyone trying to decompress B into meaningful data simply by trying many decompression algorithms. (3) Add a step between steps 1 and 2 where file B is partitioned into segments and random data inserted between the segments. (4) Instead of inserting random data segments, swap segments to create a permutation of the segments. The permutation may be determined by the password used in step 2. Until now, the US government’s default position has been: If you can’t keep data secret, at least hide it on one of 24,000 federal Websites, preferably in an incompatible or obsolete format. —Wired, July 2009. Compression Curiosities People are curious and they also like curiosities (not the same thing). It is easy to find curious facts, behavior, and objects in many areas of science and technology. In the early days of computing, for example, programmers were interested in programs that print themselves. Imagine a program in a given programming language. The source code of the program is printed on paper and can easily be read. When the program is executed, it prints its own source code. Curious! The few compression curiosities that appear here are from [curiosities 08]. We often hear the following phrase “I want it all and I want it now!” When it comes to compressing data, we all want the best compressor. So, how much can a file possibly be compressed? Lossless methods routinely compress files to less than half their size and can go down to compression ratios of about 1/8 or smaller. Lossy algorithms do much better. Is it possible to compress a file by a factor of 10,000? Now that would be a curiosity. 20 Introduction The Iterated Function Systems method (IFS, Section 7.39) can compress certain files by factors of thousands. However, such files must be especially prepared (see four examples in Section 7.39.3). Given an arbitrary data file, there is no guarantee that IFS will compress it well. The 115-byte RAR-compressed file (see Section 6.22 for RAR) found at [curio.test 08] swells, when decompressed, to 4,884,863 bytes, a compression factor of 42,477! The hexadecimal listing of this (obviously contrived) file is 526172211A07003BD07308000D00000000000000CB5C74C0802800300000007F894A0002F4EE5 A3DA39B29351D35080020000000746573742E747874A1182FD22D77B8617EF782D70000000000 00000000000000000000000000000093DE30369CFB76A800BF8867F6A9FFD4C43D7B00400700 Even if we do not obtain good compression, we certainly don’t expect a compression program to increase the size of a file, but this is precisely what happens sometimes (even often). Expansion is the bane of compression and it is easy to choose the wrong data file or the wrong compressor for a given file and so end up with an expanded file (and high blood pressure into the bargain). If we try to compress a text file with a program designed specifically for compressing images or audio, often the result is expansion, even considerable expansion. Trying to compress random data very often results in expansion. The Ring was cut from Sauron’s hand by Isildur at the slopes of Mount Doom, and he in turn lost it in the River Anduin just before he was killed in an Orc ambush. Since it indirectly caused Isildur’s death by slipping from his finger, it was known in Gondorian lore as Isildur’s Bane. —From Wikipedia Another contrived file is the 99,058-byte strange.rar, located at [curio.strange 08]. When decompressed with RAR, the resulting file, Compression Test.gba, is 1,048,576- bytes long, indicating a healthy compression factor of 10.59. However, when Compression Test.gba is compressed in turn with zip, there is virtually no compression. Two high-performance compression algorithms yield very different results when processing the same data. Curious behavior! Can a compression algorithm compress a file to itself? Obviously, the trivial algorithm that does nothing, achieves precisely that. Thus, we better ask, given a wellknown, nontrivial compression algorithm, is it possible to find (or construct) a file that the algorithm will compress to itself? The surprising answer is yes. File selfgz.gz that can be found at http://www.maximumcompression.com/selfgz.gz yields, when compressed by gzip, itself! “Curiouser and curiouser!” Cried Alice (she was so much surprised, that for the moment she quite forgot how to speak good English). —Lewis Carroll, Alice’s Adventures in Wonderland (1865) Introduction 21 The Ten Commandments of Compression 1. Redundancy is your enemy, eliminate it. 2. Entropy is your goal, strive to achieve it. 3. Read the literature before you try to publish/implement your new, clever compression algorithm. Others may have been there before you. 4. There is no universal compression method that can compress any file to just a few bytes. Thus, refrain from making incredible claims. They will come back to haunt you. 5. The G-d of compression prefers free and open source codecs. 6. If you decide to patent your algorithm, make sure anyone can understand your patent application. Others might want to improve on it. Talking about patents, recall the following warning about them (from D. Knuth) “there are better ways to earn a living than to prevent other people from making use of one’s contributions to computer science.” 7. Have you discovered a new, universal, simple, fast, and efficient compression algorithm? Please don’t ask others (especially these authors) to evaluate it for you for free. 8. The well-known saying (also from Knuth) “beware of bugs in the above code; I have only proved it correct, not tried it,” applies also (perhaps even mostly) to compression methods. Implement and test your method thoroughly before you brag about it. 9. Don’t try to optimize your algorithm/code to squeeze the last few bits out of the output in order to win a prize. Instead, identify the point where you start getting diminishing returns and stop there. 10. This is your own, private commandment. Grab a pencil, write it here, and obey it. Why this book? Most drivers know little or nothing about the operation of the engine or transmission in their cars. Few know how cellular telephones, microwave ovens, or combination locks work. Why not let scientists develop and implement compression methods and have us use them without worrying about the details? The answer, naturally, is curiosity. Many drivers try to tinker with their car out of curiosity. Many weekend sailors love to mess about with boats even on weekdays, and many children spend hours taking apart a machine, a device, or a toy in an attempt to understand its operation. If you are curious about data compression, this book is for you. The typical reader of this book should have a basic knowledge of computer science; should know something about programming and data structures; feel comfortable with terms such as bit, mega, ASCII, file, I/O, and binary search; and should be curious. The necessary mathematical background is minimal and is limited to logarithms, matrices, polynomials, differentiation/integration, and the concept of probability. This book is not intended to be a guide to software implementors and has few programs. The following URLs have useful links and pointers to the many data compression resources available on the Internet and elsewhere: http://www.hn.is.uec.ac.jp/~arimura/compression_links.html, 22 Introduction http://cise.edu.mie-u.ac.jp/~okumura/compression.html, http://compression-links.info/, http://compression.ca/ (mostly comparisons), http://datacompression.info/. This URL has a wealth of information on data compression, including tutorials, links, and lists of books. The site is owned by Mark Nelson. http://directory.google.com/Top/Computers/Algorithms/Compression/ is also a growing, up-to-date, site. Reference [Okumura 98] discusses the history of data compression in Japan. Data Compression Resources A vast number of resources on data compression are available. Any Internet search under “data compression,” “lossless data compression,” “image compression,” “audio compression,” and similar topics returns at least tens of thousands of results. Traditional (printed) resources range from general texts and texts on specific aspects or particular methods, to survey articles in magazines, to technical reports and research papers in scientific journals. Following is a short list of (mostly general) books, sorted by date of publication. Khalid Sayood, Introduction to Data Compression, Morgan Kaufmann, 3rd edition (2005). Ida Mengyi Pu, Fundamental Data Compression, Butterworth-Heinemann (2005). Darrel Hankerson, Introduction to Information Theory and Data Compression, Chapman & Hall (CRC), 2nd edition (2003). Peter Symes, Digital Video Compression, McGraw-Hill/TAB Electronics (2003). Charles Poynton, Digital Video and HDTV Algorithms and Interfaces, Morgan Kaufmann (2003). Iain E. G. Richardson, H.264 and MPEG-4 Video Compression: Video Coding for Next Generation Multimedia, John Wiley and Sons (2003). Khalid Sayood, Lossless Compression Handbook, Academic Press (2002). Touradj Ebrahimi and Fernando Pereira, The MPEG-4 Book, Prentice Hall (2002). Adam Drozdek, Elements of Data Compression, Course Technology (2001). David Taubman and Michael Marcellin, (eds), JPEG2000: Image Compression Fundamentals, Standards and Practice, Springer Verlag (2001). Kamisetty R. Rao, The Transform and Data Compression Handbook, CRC (2000). Ian H. Witten, Alistair Moffat, and Timothy C. Bell, Managing Gigabytes: Compressing and Indexing Documents and Images, Morgan Kaufmann, 2nd edition (1999). PeterWayner, Compression Algorithms for Real Programmers, Morgan Kaufmann (1999). John Miano, Compressed Image File Formats: JPEG, PNG, GIF, XBM, BMP, ACM Press and Addison-Wesley Professional (1999). Mark Nelson and Jean-Loup Gailly, The Data Compression Book, M&T Books, 2nd edition (1995). William B. Pennebaker and Joan L. Mitchell, JPEG: Still Image Data Compression Standard, Springer Verlag (1992). Timothy C. Bell, John G. Cleary, and Ian H. Witten, Text Compression, Prentice Hall (1990). James A. Storer, Data Compression: Methods and Theory, Computer Science Press. John Woods, ed., Subband Coding, Kluwer Academic Press (1990). Introduction 23 Notation The symbol “” is used to indicate a blank space in places where spaces may lead to ambiguity. The acronyms MSB and LSB refer to most-significant-bit and least-significant-bit, respectively. The notation 1i0j indicates a bit string of i consecutive 1’s followed by j zeros. Some readers called into question the title of the predecessors of this book. What does it mean for a work of this kind to be complete, and how complete is this book? Here is our opinion on the matter. We like to believe that if the entire field of data compression were (heaven forbid) to disappear, a substantial part of it could be reconstructed from this work. Naturally, we don’t compare ourselves to James Joyce, but his works provide us with a similar example. He liked to claim that if the Dublin of his time were to be destroyed, it could be reconstructed from his works. Readers who would like to get an idea of the effort it took to write this book should consult the Colophon. The authors welcome any comments, suggestions, and corrections. They should be sent to dsalomon@csun.edu or gim@ieee.org. The days just prior to marriage are like a snappy introduction to a tedious book. —Wilson Mizner 1 Basic Techniques 1.1 Intuitive Compression Data is compressed by reducing its redundancy, but this also makes the data less reliable, more prone to errors. Increasing the integrity of data, on the other hand, is done by adding check bits and parity bits, a process that increases the size of the data, thereby increasing redundancy. Data compression and data reliability are therefore opposites, and it is interesting to note that the latter is a relatively recent field, whereas the former existed even before the advent of computers. The sympathetic telegraph, discussed in the Preface, the Braille code of 1820 (Section 1.1.1), the shutter telegraph of the British admiralty [Bell et al. 90], and the Morse code of 1838 (Table 2.1) use simple, intuitive forms of compression. It therefore seems that reducing redundancy comes naturally to anyone who works on codes, but increasing it is something that “goes against the grain” in humans. This section discusses simple, intuitive compression methods that have been used in the past. Today these methods are mostly of historical interest, since they are generally inefficient and cannot compete with the modern compression methods developed during the last several decades. 1.1.1 Braille This well-known code, which enables the blind to read, was developed by Louis Braille in the 1820s and is still in common use today, after having been modified several times. Many books in Braille are available from the National Braille Press. The Braille code consists of groups (or cells) of 3Χ2 dots each, embossed on thick paper. Each of the six dots in a group may be flat or raised, implying that the information content of a group is equivalent to six bits, resulting in 64 possible groups. The letters (Table 1.1), digits, and common punctuation marks do not require all 64 codes, which is why the remaining DOI 10.1007/978-1-84882-903-9_1, © Springer-Verlag London Limited 2010 D. Salomon, G. Motta, Handbook of Data Compression, 5th ed. 26 1. Basic Techniques groups may be used to code common words—such as and, for, and of—and common strings of letters—such as ound, ation and th (Table 1.2). A •.. ... B • . • . .. C •.. •.. D •. •. •. E • . .. •. F •••. .. G ••. ••. H • . •. •. I . ••. .. J . •. ••. K • . .. • . L • . • . • . M •. •. • . N •. •••. O • . .••. P •••. • . Q •••• • . R • . •••. S . ••. • . T . ••• • . U • . .. •• V • . • . •• W. •. ••• X •. •. •• Y •. •••• Z • . .••• Table 1.1: The 26 Braille Letters. and •••. •• for •••• •• of • . •••• the . ••. •• with . ••• •• ch • . .. . • gh • . • . .• sh •.. •.• th •. •. •• Table 1.2: Some Words and Strings in Braille. Redundancy in Everyday Situations Even though we don’t unnecessarily increase redundancy in our data, we use redundant data all the time, mostly without noticing it. Here are some examples: All natural languages are redundant. A Portuguese who does not speak Italian may read an Italian newspaper and still understand most of the news because he recognizes the basic form of many Italian verbs and nouns and because most of the text he does not understand is superfluous (i.e., redundant). PIN is an acronym for “Personal Identification Number,” but banks always ask you for your “PIN number.” SALT was an acronym for “Strategic Arms Limitations Talks,” but TV announcers in the 1970s kept talking about the “SALT Talks.” These are just two examples illustrating how natural it is to be redundant in everyday situations. More examples can be found at URL http://www.corsinet.com/braincandy/twice.html  Exercise 1.1: Find a few more everyday redundant phrases. The amount of compression achieved by Braille is small but significant, because books in Braille tend to be very large (a single group covers the area of about ten printed letters). Even this modest compression comes with a price. If a Braille book is mishandled or becomes old and some dots are flattened, serious reading errors may result since every possible group is used. (Windots2, from [windots 06] and Duxbury Braille Translator, from [afb 06], are current programs for those wanting to experiment with Braille.) 1.1.2 Irreversible Text Compression Sometimes it is acceptable to “compress” text by simply throwing away some information. This is called irreversible text compression or compaction. The decompressed text will not be identical to the original, so such methods are not general purpose; they can only be used in special cases. 1.1 Intuitive Compression 27 A run of consecutive blank spaces may be replaced by a single space. This may be acceptable in literary texts and in most computer programs, but it should not be used when the data is in tabular form. In extreme cases all text characters except letters and spaces may be thrown away, and the letters may be case flattened (converted to all lower- or all uppercase). This will leave just 27 symbols, so a symbol can be encoded in 5 instead of the usual 8 bits. The compression ratio is 5/8 = .625, not bad, but the loss may normally be too great. (An interesting example of similar text is the last chapter of Ulysses by James Joyce. In addition to letters, digits, and spaces, this long chapter contains only a few punctuation marks.)  Exercise 1.2: A character set including the 26 uppercase letters and the space can be coded with 5-bit codes, but that would leave five unused codes. Suggest a way to use them. 1.1.3 Irreversible Compression The following method achieves unheard-of compression factors. We consider the input data a stream of bytes, read it byte by byte, and count the number of times each byte value occurs. There are 256 byte values, so we end up with 256 integers. After being encoded with variable-length codes, these integers become the output file. The resulting compression factor is astounding, especially if the file is large. For example, a 4 Gbyte file (about the size of a DVD) may be compressed by a factor of 650,000! Here is why. In one extreme case, all 256 counts would be roughly equal, so each would equal approximately 4G/256 = 22 Χ 230/28 = 224 and each would require about 24 bits. The total compressed size would be about 24 Χ 256 = 6,144 bits, yielding a compression factor of 4 Χ 230/6,144 . 651,000. In the other extreme case, all the bytes in the input would be identical. One count would become 4G (a 32-bit number), and the other 255 counts would be zeros (and would require one or two bits each to encode). The total number of bits would be 32 + 255 Χ 2 = 543, leading to another fantastic compression factor. With such an excellent algorithm, why do researchers still attempt to develop other methods? Answer, because the compressed file cannot be decompressed; the method is irreversible. However, we live in such a complex world that even this seemingly useless scheme may find applications. Here is an interesting example. Today, when a scientist discovers or develops something new, they rush to publish it, but in the past scientists often kept their discoveries secret for various reasons. The great mathematician Gauss did not consider much of his work worth publishing. Isaac Newton did not publish his work on the calculus until Gottfried Leibniz came up with his version. Charles Darwin kept his theory of evolution private for fear of violent reaction by the church. Niccolo Tartaglia discovered how to solve cubic equations but kept his method secret in order to win a competition (in 1535). So what do you do if you don’t want to publish a discovery, but still want to be able to claim priority in case someone else discovers the same thing? Simple; you write down a description of your achievement, compress it with the method above, keep the original in a safe place, and give the compressed version to a third party, perhaps a lawyer. When someone else claims the discovery, you show your description, compress 28 1. Basic Techniques it in the presence of witnesses, and show that the result is identical to what the third party has received from you in the past. Clever! 1.1.4 Ad Hoc Text Compression Here are some simple, intuitive ideas for cases where the compression must be reversible (lossless). If the text contains many spaces but they are not clustered, they may be removed and their positions indicated by a bit-string that contains a 0 for each text character that is not a space and a 1 for each space. Thus, the text Herearesomeideas is encoded as the bit-string 0000100010000100000 followed by the text Herearesomeideas. If the number of blank spaces is small, the bit-string will be sparse, and the methods of Section 11.5 can be employed to compress it efficiently. Since ASCII codes are essentially 7 bits long, the text may be compressed by writing 7 bits per character instead of 8 on the output stream. This may be called packing. The compression ratio is, of course, 7/8 = 0.875. The numbers 403 = 64,000 and 216 = 65,536 are not very different and satisfy the relation 403 < 216. This can serve as the basis of an intuitive compression method for a small set of symbols. If the data to be compressed is text with at most 40 different characters (such as the 26 letters, 10 digits, a space, and three punctuation marks), then this method produces a compression factor of 24/16 = 1.5. Here is how it works. Given a set of 40 characters and a string of characters from the set, we group the characters into triplets. Each character can take one of 40 values, so a trio of characters can have one of 403 = 64,000 values. Such values can be expressed in 16 bits each, because 403 is less than 216. Without compression, each of the 40 characters requires one byte, so our intuitive method produces a compression factor of 3/2 = 1.5. (This is one of those rare cases where the compression factor is constant and is known in advance.) If the text includes just uppercase letters, digits, and some punctuation marks, the old 6-bit CDC display code (Table 1.3) may be used. This code was commonly used in second-generation computers (and even a few third-generation ones). These computers did not need more than 64 characters because they did not have any display monitors and they sent their output to printers that could print only a limited set of characters. Another old code worth mentioning is the Baudot code (Table 1.4). This was a 5-bit code developed by J. M. E. Baudot in about 1880 for telegraph communication. It became popular and by 1950 was designated the International Telegraph Code No. 1. It was used in many first- and second-generation computers. The code uses 5 bits per character but encodes more than 32 characters. Each 5-bit code can be the code of two characters, a letter and a figure. The “letter shift” and “figure shift” codes are used to shift between letters and figures. Using this technique, the Baudot code can represent 32 Χ 2 - 2 = 62 characters (each code can have two meanings except the LS and FS codes). The actual number of 1.1 Intuitive Compression 29 Bits Bit positions 210 543 0 1 2 3 4 5 6 7 0 A B C D E F G 1 H I J K L M N O 2 P Q R S T U V W 3 X Y Z 0 1 2 3 4 4 5 6 7 8 9 + - * 5 / ( ) $ = sp , . 6 . [ ] : = — . . 7 ^ v < > . . ¬ ; Table 1.3: The CDC Display Code. characters, however, is smaller than that, because five of the codes have one meaning each, and some codes are not assigned. The Baudot code is not reliable because it does not employ a parity bit. A bad bit transforms a character into another character. In particular, a bad bit in a shift character causes a wrong interpretation of all the characters following, up to the next shift. Letters Code Figures Letters Code Figures A 10000 1 Q 10111 / B 00110 8 R 00111 - C 10110 9 S 00101 SP D 11110 0 T 10101 na E 01000 2 U 10100 4 F 01110 na V 11101 ' G 01010 7 W 01101 ? H 11010 + X 01001 , I 01100 na Y 00100 3 J 10010 6 Z 11001 : K 10011 ( LS 00001 LS L 11011 = FS 00010 FS M 01011 ) CR 11000 CR N 01111 na LF 10001 LF O 11100 5 ER 00011 ER P 11111 % na 00000 na LS, Letter Shift; FS, Figure Shift. CR, Carriage Return; LF, Line Feed. ER, Error; na, Not Assigned; SP, Space. Table 1.4: The Baudot Code. If the data includes just integers, each decimal digit may be represented in 4 bits, with two digits packed in a byte. Data consisting of dates may be represented as the 30 1. Basic Techniques number of days since January 1, 1900 (or some other convenient start date). Each date may be stored as a 16-bit or 24-bit number (2 or 3 bytes). If the data consists of date/time pairs, a possible compressed representation is the number of seconds since a convenient start date. If stored as a 32-bit number (4 bytes) such a representation can be sufficient for about 136 years. Dictionary data (or any list sorted lexicographically) can be compressed using the concept of front compression. This is based on the observation that adjacent words in such a list tend to share some of their initial characters. A word can therefore be compressed by dropping the n characters it shares with its predecessor in the list and replacing them with the numeric value of n. The 9/19/89 Syndrome How can a date, such as 11/12/71, be represented inside a computer? One way to do this is to store the number of days since January 1, 1900 in an integer variable. If the variable is 16 bits long (including 15 magnitude bits and one sign bit), it will overflow after 215 = 32K = 32,768 days, which is September 19, 1989. This is precisely what happened on that day in several computers (see the January, 1991 issue of the Communications of the ACM). Notice that doubling the size of such a variable to 32 bits would have delayed the problem until after 231 = 2 giga days have passed, which would occur sometime in the fall of year 5,885,416. Table 1.5 shows a short example taken from a word list used to create anagrams. It is clear that it is easy to obtain significant compression with this simple method (see also [Robinson and Singer 81] and [Nix 81]). a a aardvark 1ardvark aback 1back abaft 3ft abandon 3ndon abandoning 7ing abasement 3sement abandonment 3ndonment abash 3sh abated 3ted abate 5 abbot 2bot abbey 3ey abbreviating 3reviating abbreviate 9e abbreviation 9ion Table 1.5: Front Compression. 1.2 Run-Length Encoding 31 The MacWrite word processor [Young 85] used a special 4-bit code to encode the most-common 15 characters “etnroaisdlhcfp” plus an escape code. Any of these 15 characters is encoded by 4 bits. Any other character is encoded as the escape code followed by the 8 bits of the ASCII code of the character; a total of 12 bits. Each paragraph is coded separately, and if this results in expansion, the paragraph is stored as plain ASCII. One more bit is added to each paragraph to indicate whether or not it uses compression. The principle of parsimony values a theory’s ability to compress a maximum of information into a minimum of formalism. Einstein’s celebrated E = mc2 derives part of its well-deserved fame from the astonishing wealth of meaning it packs into its tiny frame. Maxwell’s equations, the rules of quantum mechanics, and even the basic equations of the general theory of relativity similarly satisfy the parsimony requirement of a fundamental theory: They are compact enough to fit on a T-shirt. By way of contrast, the human genome project, requiring the quantification of hundreds of thousands of genetic sequences, represents the very antithesis of parsimony. —Hans C. von Baeyer, Maxwell’s Demon, 1998 1.2 Run-Length Encoding The idea behind this approach to data compression is this: If a data item d occurs n consecutive times in the input stream, replace the n occurrences with the single pair nd. The n consecutive occurrences of a data item are called a run length of n, and this approach to data compression is called run-length encoding or RLE. We apply this idea first to text compression and then to image compression. 1.3 RLE Text Compression Just replacing 2.allistoowell with 2.a2ist2we2 will not work. Clearly, the decompressor should have a way to tell that the first 2 is part of the text while the others are repetition factors for the letters o and l. Even the string 2.a2list2owe2l does not solve this problem (and also does not produce any compression). One way to solve this problem is to precede each repetition with a special escape character. If we use the character @ as the escape character, then the string 2.a@2list@2owe@2l can be decompressed unambiguously. However, this string is longer than the original string, because it replaces two consecutive letters with three characters. We have to adopt the convention that only three or more repetitions of the same character will be replaced with a repetition factor. Figure 1.6a is a flowchart for such a simple run-length text compressor. After reading the first character, the repeat-count is 1 and the character is saved. Subsequent characters are compared with the one already saved, and if they are identical 32 1. Basic Techniques to it, the repeat-count is incremented. When a different character is read, the operation depends on the value of the repeat count. If it is small, the saved character is written on the compressed file and the newly-read character is saved. Otherwise, an @ is written, followed by the repeat-count and the saved character. Decompression is also straightforward. It is shown in Figure 1.6b. When an @ is read, the repetition count n and the actual character are immediately read, and the character is written n times on the output stream. The main problems with this method are the following: 1. In plain English text there are not many repetitions. There are many “doubles” but a “triple” is rare. The most repetitive character is the space. Dashes or asterisks may sometimes also repeat. In mathematical texts, digits may repeat. The following “paragraph” is a contrived example. The abbott from Abruzzi accedes to the demands of all abbesses from Narragansett and Abbevilles from Abyssinia. He will accommodate them, abbreviate his sabbatical, and be an accomplished accessory. 2. The character “@” may be part of the text in the input stream, in which case a different escape character must be chosen. Sometimes the input stream may contain every possible character in the alphabet. An example is an object file, the result of compiling a program. Such a file contains machine instructions and can be considered a string of bytes that may have any values. The MNP5 method described below and in Section 5.4 provides a solution. 3. Since the repetition count is written on the output stream as a byte, it is limited to counts of up to 255. This limitation can be softened somewhat when we realize that the existence of a repetition count means that there is a repetition (at least three identical consecutive characters). We may adopt the convention that a repeat count of 0 means three repeat characters, which implies that a repeat count of 255 means a run of 258 identical characters. There are three kinds of lies: lies, damned lies, and statistics. (Attributed by Mark Twain to Benjamin Disraeli) The MNP class 5 method was used for data compression in old modems. It has been developed by Microcom, Inc., a maker of modems (MNP stands for Microcom Network Protocol), and it uses a combination of run-length and adaptive frequency encoding. The latter technique is described in Section 5.4, but here is how MNP5 solves problem 2 above. When three or more identical consecutive bytes are found in the input stream, the compressor writes three copies of the byte on the output stream, followed by a repetition count. When the decompressor reads three identical consecutive bytes, it knows that the next byte is a repetition count (which may be 0, indicating just three repetitions). A disadvantage of the method is that a run of three characters in the input stream results in four characters written to the output stream; expansion! A run of four characters results in no compression. Only runs longer than four characters get compressed. Another slight 1.3 RLE Text Compression 33 SC:=save CH R:=R+1 Start 1 1 1 (a) Yes No Yes No Yes No Stop Yes No Char. count C:=0 Repeat count R:=0 Read next character, CH C:=C+1 C=1? Write SC on output file R times Write compressed format (3 chars) R:=0 goto 1. SC:=Save CH SC=CH R<4 eof? Figure 1.6: RLE. Part I: Compression. 34 1. Basic Techniques (b) Read n. Read n Chars. Generate n repetitions Compression flag:=on Compression flag off? fl Start yes no Stop Write char on output yes yes no no Compression flag:=off Read next character eof? Char=‘@’ Figure 1.6 RLE. Part II: Decompression. 1.3 RLE Text Compression 35 problem is that the maximum count is artificially limited in MNP5 to 250 instead of 255. To get an idea of the compression ratios produced by RLE, we assume a string of N characters that needs to be compressed. We assume that the string contains M repetitions of average length L each. Each of the M repetitions is replaced by 3 characters (escape, count, and data), so the size of the compressed string is N - M Χ L +M Χ3 = N -M(L - 3) and the compression factor is N N -M(L - 3) . (For MNP5 just substitute 4 for 3.) Examples: N = 1000,M = 10, L = 3 yield a compression factor of 1000/[1000 - 10(4 - 3)] = 1.01. A better result is obtained in the case N = 1000,M = 50, L = 10, where the factor is 1000/[1000 - 50(10 - 3)] = 1.538. A variant of run-length encoding for text is digram encoding. This method is suitable for cases where the data to be compressed consists only of certain characters, e.g., just letters, digits, and punctuation marks. The idea is to identify commonly-occurring pairs of characters and to replace a pair (a digram) with one of the characters that cannot occur in the data (e.g., one of the ASCII control characters). Good results can be obtained if the data can be analyzed beforehand. We know that in plain English certain pairs of characters, such as E, T, TH, and A, occur often. Other types of data may have different common digrams. The sequitur method of Section 11.10 is an example of a method that compresses data by locating repeating digrams (as well as longer repeated phrases) and replacing them with special symbols. A similar variant is pattern substitution. This is suitable for compressing computer programs, where certain words, such as for, repeat, and print, occur often. Each such word is replaced with a control character or, if there are many such words, with an escape character followed by a code character. Assuming that code a is assigned to the word print, the text m:print,b,a; will be compressed to m:@a,b,a;. 1.3.1 Relative Encoding This is another variant, sometimes called differencing (see [Gottlieb et al. 75]). It is used in cases where the data to be compressed consists of a string of numbers that do not differ by much, or in cases where it consists of strings that are similar to each other. An example of the former is telemetry. The latter case is used in facsimile data compression described in Section 5.7 and also in LZW compression (Section 6.13.4). In telemetry, a sensing device is used to collect data at certain intervals and transmit it to a central location for further processing. An example is temperature values collected every hour. Successive temperatures normally do not differ by much, so the sensor needs to send only the first temperature, followed by differences. Thus the sequence of temperatures 70, 71, 72.5, 73.1, . . . can be compressed to 70, 1, 1.5, 0.6, . . .. This compresses the data, because the differences are small and can be expressed in fewer bits. Notice that the differences can be negative and may sometimes be large. When a large difference is found, the compressor sends the actual value of the next measurement instead of the difference. Thus, the sequence 110, 115, 121, 119, 200, 202, . . . can be compressed to 110, 5, 6,-2, 200, 2, . . .. Unfortunately, we now need to distinguish between a 36 1. Basic Techniques difference and an actual value. This can be done by the compressor creating an extra bit (a flag) for each number sent, accumulating those bits, and sending them to the decompressor from time to time, as part of the transmission. Assuming that each difference is sent as a byte, the compressor should follow (or precede) a group of eight bytes with a byte consisting of their eight flags. Another practical way to send differences mixed with actual values is to send pairs of bytes. Each pair is either an actual 16-bit measurement (positive or negative) or two 8-bit signed differences. Thus, actual measurements can be between 0 and ±32K and differences can be between 0 and ±255. For each pair, the compressor creates a flag: 0 if the pair is an actual value, 1 if it is a pair of differences. After 16 pairs are sent, the compressor sends the 16 flags. Example: The sequence of measurements 110, 115, 121, 119, 200, 202, . . . is sent as (110), (5, 6), (-2,-1), (200), (2, . . .), where each pair of parentheses indicates a pair of bytes. The -1 has value 111111112, which is ignored by the decompressor (it indicates that there is only one difference in this pair). While sending this information, the compressor prepares the flags 01101 . . ., which are sent after the first 16 pairs. Relative encoding can be generalized to the lossy case, where it is called differential encoding. An example of a differential encoding method is differential pulse code modulation (or DPCM, Section 7.30). 1.4 RLE Image Compression RLE is a natural candidate for compressing graphical data. A digital image consists of small dots called pixels. (For information on pixels and their history, see Section 7.1 as well as the lively references [Lyon 09] and [Smith 09].) Each pixel can be either one bit, indicating a black or a white dot, or several bits, indicating one of several colors or shades of gray. We assume that the pixels are stored in an array called a bitmap in memory, so the bitmap is the input stream for the image. Pixels are normally arranged in the bitmap in scan lines, so the first bitmap pixel is the dot at the top-left corner of the image, and the last pixel is the one at the bottom-right corner. Compressing an image using RLE is based on the observation that if we select a pixel in the image at random, there is a good chance that its neighbors will have the same color (see also Sections 7.34 and 7.36). The compressor therefore scans the bitmap row by row, looking for runs of pixels of the same color. If the bitmap starts, e.g., with 17 white pixels, followed by 1 black pixel, followed by 55 white ones, etc., then only the numbers 17, 1, 55, . . . need be written on the output stream. The compressor assumes that the bitmap starts with white pixels. If this is not true, then the bitmap starts with zero white pixels, and the output stream should start with the run length 0. The resolution of the bitmap should also be saved at the start of the output stream. The size of the compressed stream depends on the complexity of the image. The more detail, the worse the compression. However, Figure 1.7 shows how scan lines pass through a uniform region. A line enters through one point on the perimeter of the region and exits through another point, and these two points are not “used” by any other scan lines. It is now clear that the number of scan lines traversing a uniform region is roughly 1.4 RLE Image Compression 37 equal to half the length (measured in pixels) of its perimeter. Since the region is uniform, each scan line contributes two runs to the output stream for each region it crosses. The compression ratio of a uniform region therefore roughly equals the ratio 2Χhalf the length of the perimeter total number of pixels in the region = perimeter area . Figure 1.7: Uniform Areas and Scan Lines.  Exercise 1.3: What would be the compressed file in the case of the following 6 Χ 8 bitmap? RLE can also be used to compress grayscale images. Each run of pixels of the same intensity (gray level) is encoded as a pair (run length, pixel value). The run length usually occupies one byte, allowing for runs of up to 255 pixels. The pixel value occupies several bits, depending on the number of gray levels (typically between 4 and 8 bits). Example: An 8-bit-deep grayscale bitmap that starts with 12, 12, 12, 12, 12, 12, 12, 12, 12, 35, 76, 112, 67, 87, 87, 87, 5, 5, 5, 5, 5, 5, 1, . . . is compressed into 9 ,12,35,76,112,67,3 ,87, 6 ,5,1,. . . , where the boxed numbers indicate counts. The problem is to distinguish between a byte containing a grayscale value (such as 12) and one containing a count (such as 9 ). Here are some solutions (although not the only possible ones): 1. If the image is limited to just 128 grayscales, we can devote one bit in each byte to indicate whether the byte contains a grayscale value or a count. 2. If the number of grayscales is 256, it can be reduced to 255 with one value reserved as a flag to precede every byte with a count. If the flag is, say, 255, then the sequence 38 1. Basic Techniques above becomes 255, 9, 12, 35, 76, 112, 67, 255, 3, 87, 255, 6, 5, 1, . . . . 3. Again, one bit is devoted to each byte to indicate whether the byte contains a grayscale value or a count. This time, however, these extra bits are accumulated in groups of 8, and each group is written on the output stream preceding (or following) the 8 bytes it “corresponds to.” Example: the sequence 9 ,12,35,76,112,67,3 ,87, 6 ,5,1,. . . becomes 10000010,9,12,35,76,112,67,3,87,100..... ,6,5,1,. . . . The total size of the extra bytes is, of course, 1/8 the size of the output stream (they contain one bit for each byte of the output stream), so they increase the size of that stream by 12.5%. 4. A group of m pixels that are all different is preceded by a byte with the negative value -m. The sequence above is encoded by 9, 12,-4, 35, 76, 112, 67, 3, 87, 6, 5, ?, 1, . . . (the value of the byte with ? is positive or negative depending on what follows the pixel of 1). The worst case is a sequence of pixels (p1, p2, p2) repeated n times throughout the bitmap. It is encoded as (-1, p1, 2, p2), four numbers instead of the original three! If each pixel requires one byte, then the original three bytes are expanded into four bytes. If each pixel requires three bytes, then the original three pixels (which constitute nine bytes) are compressed into 1+3+1+3 = 8 bytes. Three more points should be mentioned: 1. Since the run length cannot be 0, it makes sense to write the [run length minus one] on the output stream. Thus, the pair (3, 87) denotes a run of four pixels with intensity 87. This way, a run can be up to 256 pixels long. 2. In color images it is common to have each pixel stored as three bytes, representing the intensities of the red, green, and blue components of the pixel. In such a case, runs of each color should be encoded separately. Thus the pixels (171, 85, 34), (172, 85, 35), (172, 85, 30), and (173, 85, 33) should be separated into the three sequences (171, 172, 172, 173, . . .), (85, 85, 85, 85, . . .), and (34, 35, 30, 33, . . .). Each sequence should be run-length encoded separately. This means that any method for compressing grayscale images can be applied to color images as well. 3. It is preferable to encode each row of the bitmap individually. Thus if a row ends with four pixels of intensity 87 and the following row starts with 9 such pixels, it is better to write . . . , 4, 87, 9, 87, . . . on the output stream rather than . . . , 13, 87, . . .. It is even better to write the sequence . . . , 4, 87, eol, 9, 87, . . ., where “eol” is a special end-of-line code. The reason is that sometimes the user may decide to accept or reject an image just by examining its general shape, without any details. If each line is encoded individually, the decoding algorithm can start by decoding and displaying lines 1, 6, 11, . . ., continue with lines 2, 7, 12, . . ., etc. The individual rows of the image are interlaced, and the image is constructed on the screen gradually, in several steps. This way, it is possible to get an idea of what is in the image at an early stage. Figure 1.8c shows an example of such a scan. 1.4 RLE Image Compression 39 1 2 3 4 5 6 7 8 9 10 (a) (b) (c) Figure 1.8: RLE Scanning. Another advantage of individual encoding of rows is to make it possible to extract just part of an encoded image (such as rows k through l). Yet another application is to merge two compressed images without having to decompress them first. If this idea (encoding each bitmap row individually) is adopted, then the compressed stream must contain information on where each bitmap row starts in the stream. This can be done by writing a header at the start of the stream that contains a group of 4 bytes (32 bits) for each bitmap row. The kth such group contains the offset (in bytes) from the start of the stream to the start of the information for row k. This increases the size of the compressed stream but may still offer a good trade-off between space (size of compressed stream) and time (time to decide whether to accept or reject the image).  Exercise 1.4: There is another, obvious, reason why each bitmap row should be coded individually. What is it? Figure 1.9a lists Matlab code to compute run lengths for a bi-level image. The code is very simple. It starts by flattening the matrix into a one-dimensional vector, so the run lengths continue from row to row. Disadvantage of image RLE: When the image is modified, the run lengths normally have to be completely redone. The RLE output can sometimes be bigger than pixelby- pixel storage (i.e., an uncompressed image, a raw dump of the bitmap) for complex pictures. Imagine a picture with many vertical lines. When it is scanned horizontally, it produces very short runs, resulting in very bad compression, or even in expansion. A good, practical RLE image compressor should be able to scan the bitmap by rows, columns, or in zigzag (Figure 1.8a,b) and it may automatically try all three ways on every bitmap to achieve the best compression.  Exercise 1.5: Given the 8Χ8 bitmap of Figure 1.9b, use RLE to compress it, first row by row, then column by column. Describe the results in detail. 40 1. Basic Techniques % Returns the run lengths of % a matrix of 0s and 1s function R=runlengths(M) [c,r]=size(M); for i=1:c; x(r*(i-1)+1:r*i)=M(i,:); end N=r*c; y=x(2:N); u=x(1:N-1); z=y+u; j=find(z==1); i1=[j N]; i2=[0 j]; R=i1-i2; the test M=[0 0 0 1; 1 1 1 0; 1 1 1 0] runlengths(M) produces 3 4 1 3 1 (a) (b) Figure 1.9: (a) Matlab Code To Compute Run Lengths. (b) A Bitmap. 1.4.1 Lossy Image Compression It is possible to get even better compression if short runs are ignored. Such a method loses information when compressing an image, but sometimes this is acceptable to the user. (Images that permit no loss are medical X-rays and pictures taken by large telescopes, where the price of an image is astronomical.) A lossy run-length encoding algorithm should start by asking the user for the longest run that should still be ignored. If the user specifies, for example, 3, then the program merges all runs of 1, 2, or 3 identical pixels with their neighbors. The compressed data “6,8,1,2,4,3,11,2” would be saved, in this case, as “6,8,7,16” where 7 is the sum 1+2+4 (three runs merged) and 16 is the sum 3 + 11 + 2. This makes sense for large highresolution images where the loss of some detail may be invisible to the eye, but may significantly reduce the size of the output stream (see also Chapter 7). 1.4.2 Conditional Image RLE Facsimile compression (Section 5.7) uses a modified Huffman code, but it can also be considered a modified RLE. This section discusses another modification of RLE, proposed in [Gharavi 87]. Assuming an image with n grayscales, the method starts by assigning an n-bit code to each pixel depending on its near neighbors. It then concate- 1.4 RLE Image Compression 41 nates the n-bit codes into a long string and computes run lengths. The run lengths are assigned prefix codes (Huffman or other, Section 2.2) that are written on the compressed stream. The method considers each scan line in the image a second-order Markov model. In such a model the value of the current data item depends on just two of its past neighbors, not necessarily the two immediate ones. Figure 1.10 shows the two neighbors A and B used by our method to predict the current pixel X (compare this with the lossless mode of JPEG, Section 7.10.5). A set of training images is used to count—for each possible pair of values of the neighbors A, B—how many times each value of X occurs. If A and B have similar values, it is natural to expect that X will have a similar value. If A and B have very different values, we expect X to have many different values, each with a low probability. The counts therefore produce the conditional probabilities P(X|A,B) (the probability of the current pixel having value X if we already know that its neighbors have values A and B). Table 1.11 lists a small part of the results obtained by counting this way several training images with 4-bit pixels. B A X Figure 1.10: Neighbors Used To Predict X. A B W1 W2 W3 W4 W5 W6 W7 . . . 2 15 value: 4 3 10 0 6 8 1. . . count: 21 6 5 4 2 2 1. . . 3 0 value: 0 1 3 2 11 4 15 . . . count: 443 114 75 64 56 19 12 . . . 3 1 value: 1 2 3 4 0 5 6. . . count: 1139 817 522 75 55 20 8 . . . 3 2 value: 2 3 1 4 5 6 0. . . count: 7902 4636 426 264 64 18 6 . . . 3 3 value: 3 2 4 5 1 6 7. . . count: 33927 2869 2511 138 93 51 18 . . . 3 4 value: 4 3 5 2 6 7 1. . . count: 2859 2442 240 231 53 31 13 . . . Table 1.11: Conditional Counting of 4-Bit Pixels. Each pixel in the image to be compressed is assigned a new 4-bit code depending on its conditional probability as indicated by the table. Imagine a current pixel X with value 1 whose neighbors have values A = 3, B = 1. The table indicates that the conditional probability P(1|3, 1) is high, so X should be assigned a new 4-bit code with few runs (i.e., codes that contain consecutive 1’s or consecutive 0’s). On the other hand, 42 1. Basic Techniques the same X = 1 with neighbors A = 3 and B = 3 can be assigned a new 4-bit code with many runs, since the table indicates that the conditional probability P(1|3, 3) is low. The method therefore uses conditional probabilities to detect common pixels (hence the name conditional RLE), and assigns them codes with few runs. Examining all 16 four-bit codes W1 through W16, we find that the two codes 0000 and 1111 have one run each, while 0101 and 1010 have four runs each. The codes should be arranged in order of increasing runs. For 4-bit codes we end up with the four groups 1. W1 to W2 : 0000, 1111, 2. W3 to W8 : 0001, 0011, 0111, 1110, 1100, 1000, 3. W9 to W14 : 0100, 0010, 0110, 1011, 1101, 1001, 4. W15 to W16 : 0101, 1010. The codes of group i have i runs each. The codes in each group are selected such that the codes in the second half of a group are the complements of those in the first half.  Exercise 1.6: Apply this principle to construct the 32 five-bit codes. The method can now be described in detail. The image is scanned in raster order. For each pixel X, its neighbors A and B are located, and the table is searched for this triplet. If the triplet is found in column i, then code Wi is selected. The first pixel has no neighbors, so assuming that its value is i, code Wi is selected for it. If X is located at the top row of the image, it has an A neighbor but not a B neighbor, so this case is handled differently. Imagine a pixel X = 2 at the top row with a neighbor A = 3. All the rows in the table with A = 3 are examined, in this case, and the one with the largest count for X = 2 is selected. In our example this is the row with count 7902, so code W1 is selected. Pixels X with no A neighbors (on the left column) are treated similarly. Rule of complementing: After the code for X has been selected, it is compared with the preceding code. If the least-significant bit of the preceding code is 1, the current code is complemented. This is supposed to reduce the number of runs. As an example, consider the typical sequence of 4-bit codes W2, W4, W1, W6, W3, W2, W1 1111, 0011, 0000, 1110, 0001, 1111, 0000. When these codes are concatenated, the resulting 28-bit string has eight runs. After applying the rule above, the codes become 1111, 1100, 0000, 1110, 0001, 0000, 0000, a string with just six runs.  Exercise 1.7: Do the same for the code sequence W1, W2, W3, W6, W1, W4, W2. A variation of this method uses the counts of Table 1.11 but not its codes and run lengths. Instead, it assigns a prefix code to the current pixel X depending on its neighbors A and B. Table 1.12 is an example. Each row has a different set of prefix codes constructed according to the counts of the row. 1.4 RLE Image Compression 43 A B W1 W2 W3 W4 W5 W6 W7 . . . 2 15 value: 4 3 10 0 6 8 1 . . . code: 01 00 111 110 1011 1010 10010 . . . 3 0 value: 0 1 3 2 11 4 15 . . . code: 11 10 00 010 0110 011111 011101 . . . 3 1 value: 1 2 3 4 0 5 6 . . . code: 0 11 100 1011 10100 101010 10101111 . . . 3 2 value: 2 3 1 4 5 6 0 . . . code: 0 11 100 1011 10100 101010 10101111 . . . 3 3 value: 3 2 4 5 1 6 7 . . . code: 0 11 100 1011 101001 1010000 10101000 . . . 3 4 value: 4 3 5 2 6 7 1 . . . code: 11 10 00 0111 0110 0100 010110 . . . Table 1.12: Prefix Codes For 4-Bit Pixels. 1.4.3 The BinHex 4.0 Format BinHex 4.0 is an old file format for reliable file transfers, designed by Yves Lempereur for use on the Macintosh platform. Before delving into the details of the format, the reader should understand why such a format is useful. ASCII is a 7-bit code. Each character is coded as a 7-bit number, which allows for 128 characters in the ASCII table. The ASCII standard recommends adding an eighth bit as parity to every character for increased reliability. However, the standard does not specify odd or even parity, and many computers simply ignore the extra bit or even set it to 0. As a result, when files are transferred in a computer network, some transfer programs may ignore the eighth bit and transfer just seven bits per character. This isn’t so bad when a text file is being transferred but when the file is binary, no bits should be ignored. This is why it is safer to transfer text files, rather than binary files, over computer networks. The idea of BinHex is to translate any file to a text file. The BinHex program reads an input file (text or binary) and produces an output file with the following format: 1. The comment: (ThisfilemustbeconvertedwithBinHex4.0) 2. A header including the items listed in Table 1.13. 3. The input file is then read and RLE is used as the first step. Character 9016 is used as the RLE marker, and the following examples speak for themselves: Source string Packed string 00 11 22 33 44 55 66 77 00 11 22 33 44 55 66 77 11 22 22 22 22 22 22 33 11 22 90 06 33 11 22 90 33 44 11 22 90 00 33 44 (The character 00 indicates no run.) Runs of lengths 3 thorugh 255 characters are encoded in this way. 44 1. Basic Techniques Field Size Length of FileName (1–63) byte FileName (“Length” bytes) Version byte Type long Creator long Flags (And $F800) word Length of Data Fork long Length of Resource Fork long CRC word Data Fork (“Data Length” bytes) CRC word Resource Fork (“Rsrc Length” bytes) CRC word Table 1.13: The BinHex Header.  Exercise 1.8: How is the string “11 22 90 00 33 44” encoded? 4. Encoding into 7-bit ASCII characters. The input file is considered a stream of bits. As the file is being read, it is divided into blocks of 6 bits, and each block is used as a pointer to the BinHex table below. The character that’s pointed to in this table is written on the output file. The table is !"#$%&’()*+,-012345689@ABCDEFGHIJKLMNPQRSTUVXYZ[‘abcdefhijklmpqr The output file is organized in “lines” of 64 characters each (except, perhaps, the last line). Each line is preceded and followed by a pair of colons “:”. The following is a quotation from the designer: “The characters in this table have been chosen for maximum noise protection.”  Exercise 1.9: Manually convert the string “123ABC” to BinHex. Ignore the comment and the file header. 1.4.4 BMP Image Files BMP is the native format for image files in the Microsoft Windows operating system. It has been modified several times since its inception, but has remained stable from version 3 of Windows. BMP is a palette-based graphics file format for images with 1, 2, 4, 8, 16, 24, or 32 bitplanes. It uses a simple form of RLE to compress images with 4 or 8 bitplanes. The format of a BMP file is simple. It starts with a file header that contains the two bytes BM and the file size. This is followed by an image header with the width, height, and number of bitplanes (there are two different formats for this header). Following the two headers is the color palette (that can be in one of three formats) which is followed by the image pixels, either in raw format or compressed by RLE. Detailed information on the BMP file format can be found in, for example, [Miano 99] and [Swan 93]. This section discusses the particular version of RLE used by BMP to compress pixels. 1.5 Move-to-Front Coding 45 For images with eight bitplanes, the compressed pixels are organized in pairs of bytes. The first byte of a pair is a count C, and the second byte is a pixel value P which is repeated C times. Thus, the pair 0416 0216 is expanded to the four pixels 0216 0216 0216 0216. A count of 0 acts as an escape, and its meaning depends on the byte that follows. A zero byte followed by another zero indicates end-of-line. The remainder of the current image row is filled with pixels of 00 as needed. A zero byte followed by 0116 indicates the end of the image. The remainder of the image is filled up with 00 pixels. A zero byte followed by 0216 indicates a skip to another position in the image. A 0016 0216 pair must be followed by two bytes indicating how many columns and rows to skip to reach the next nonzero pixel. Any pixels skipped are filled with zeros. A zero byte followed by a byte C greater than 2 indicates C raw pixels. Such a pair must be followed by the C pixels. Assuming a 4Χ8 image with 8-bit pixels, the following sequence 0416 0216, 0016 0416 a35b124716, 0116 f516, 0216 e716, 0016 0216 000116, 0116 9916, 0316 c116, 0016 0016, 0016 0416 08926bd716, 0016 0116 is the compressed representation of the 32 pixels 02 02 02 02 a3 5b 12 47 f5 e7 e7 00 00 00 00 00 00 00 99 c1 c1 c1 00 00 08 92 6b d7 00 00 00 00 Images with four bitplanes are compressed in a similar way but with two exceptions. The first exception is that a pair (C, P) represents a count byte and a byte of two pixel values that alternate. The pair 0516 a216, for example, is the compressed representation of the five 4-bit pixels a, 2, a, 2, and a, while the pair 0716 ff16 represents seven consecutive 4-bit pixels of f16. The second exception has to do with pairs (0,C) where C is greater than 2. Such a pair is followed by C 4-bit pixel values, packed two to a byte. The value of C is normally a multiple of 4, implying that a pair (0,C) specifies pixels to fill up an integer number of words (where a word is two bytes). If C is not a multiple of 4, the remainder of the last word is padded with zeros. Thus, 0016 0816 a35b124716 specifies eight pixels and fills up four bytes (or two words) with a35b124716, whereas 0016 0616 a35b1216 specifies six pixels and also fills up four bytes but with a35b120016. 1.5 Move-to-Front Coding The basic idea of this method [Bentley 86] is to maintain the alphabet A of symbols as a list where frequently-occurring symbols are located near the front. A symbol s is encoded as the number of symbols that precede it in this list. Thus if A=(t, h, e, s,. . . ) and the next symbol in the input stream to be encoded is the e, it will be encoded as 2, since it is preceded by two symbols. There are several possible variants to this method; the most basic of them adds one more step: After symbol s is encoded, it is moved to the front of list A. Thus, after encoding the e, the alphabet is modified to A=(e, t, h, s,. . . ). This move-to-front step reflects the expectation that once e has been read from 46 1. Basic Techniques the input stream, it will be read many more times and will, at least for a while, be a common symbol. The move-to-front method is locally adaptive, since it adapts itself to the frequencies of symbols in local areas of the input stream. The method produces good results if the input stream satisfies this expectation, i.e., if it contains concentrations of identical symbols (if the local frequency of symbols changes significantly from area to area in the input stream). We call this the concentration property. Here are two examples that illustrate the move-to-front idea. Both assume the alphabet A=(a, b, c, d, m, n, o, p). 1. The input stream abcddcbamnopponm is encoded as C = (0, 1, 2, 3, 0, 1, 2, 3, 4, 5, 6, 7, 0, 1, 2, 3) (Table 1.14a). Without the move-to-front step it is encoded as C = (0, 1, 2, 3, 3, 2, 1, 0, 4, 5, 6, 7, 7, 6, 5, 4) (Table 1.14b). Both C and C contain codes in the same range [0, 7], but the elements of C are smaller on average, since the input starts with a concentration of abcd and continues with a concentration of mnop. (The average value of C is 2.5, while that of C is 3.5.) a abcdmnop 0 b abcdmnop 1 c bacdmnop 2 d cbadmnop 3 d dcbamnop 0 c dcbamnop 1 b cdbamnop 2 a bcdamnop 3 m abcdmnop 4 n mabcdnop 5 o nmabcdop 6 p onmabcdp 7 p ponmabcd 0 o ponmabcd 1 n opnmabcd 2 m nopmabcd 3 mnopabcd (a) a abcdmnop 0 b abcdmnop 1 c abcdmnop 2 d abcdmnop 3 d abcdmnop 3 c abcdmnop 2 b abcdmnop 1 a abcdmnop 0 m abcdmnop 4 n abcdmnop 5 o abcdmnop 6 p abcdmnop 7 p abcdmnop 7 o abcdmnop 6 n abcdmnop 5 m abcdmnop 4 (b) a abcdmnop 0 b abcdmnop 1 c bacdmnop 2 d cbadmnop 3 m dcbamnop 4 n mdcbanop 5 o nmdcbaop 6 p onmdcbap 7 a ponmdcba 7 b aponmdcb 7 c baponmdc 7 d cbaponmd 7 m dcbaponm 7 n mdcbapon 7 o nmdcbapo 7 p onmdcbap 7 ponmdcba (c) a abcdmnop 0 b abcdmnop 1 c abcdmnop 2 d abcdmnop 3 m abcdmnop 4 n abcdmnop 5 o abcdmnop 6 p abcdmnop 7 a abcdmnop 0 b abcdmnop 1 c abcdmnop 2 d abcdmnop 3 m abcdmnop 4 n abcdmnop 5 o abcdmnop 6 p abcdmnop 7 (d) Table 1.14: Encoding With and Without Move-to-Front. 2. The input stream abcdmnopabcdmnop is encoded as C = (0, 1, 2, 3, 4, 5, 6, 7, 7, 7, 7, 7, 7, 7, 7, 7) (Table 1.14c). Without the move-to-front step it is encoded as C = (0, 1, 2, 3, 4, 5, 6, 7, 0, 1, 2, 3, 4, 5, 6, 7) (Table 1.14d). The average of C is now 5.25, greater than that of C, which is 3.5. The move-to-front rule creates a worse result in this case, because the input does not contain concentrations of identical symbols (it does not satisfy the concentration property). Before getting into further details, it is important to understand the advantage of having small numbers in C. This feature makes it possible to efficiently encode C with either Huffman or arithmetic coding (Chapter 5). Here are four ways to do this: 1.5 Move-to-Front Coding 47 1. Assign Huffman codes to the integers in the range [0, n] such that the smaller integers get the shorter codes. Here is an example of such a code for the integers 0 through 7: 0—0, 1—10, 2—110, 3—1110, 4—11110, 5—111110, 6—1111110, 7—1111111. 2. Assign codes to the integers such that the code of integer i . 1 is its binary code preceded by log2 i zeros. This is the well-known Elias gamma code (Section 3.4). Table 1.15 lists some examples. i Code Size 1 1 1 2 010 3 3 011 3 4 00100 5 5 00101 5 6 00110 5 7 00111 5 8 0001000 7 9 0001001 7 ... ... ... 15 0001111 7 16 000010000 9 Table 1.15: Examples of Variable-Length Codes.  Exercise 1.10: What is the total size of the code of i in this case. 3. Use adaptive Huffman coding (Section 5.3). 4. For maximum compression, perform two passes over C, the first pass counts frequencies of codes and the second pass performs the actual encoding. The frequencies counted in pass 1 are used to compute probabilities and assign Huffman codes to be used later by pass 2. It can be shown that the move-to-front method performs, in the worst case, slightly worse than Huffman coding. At best, it performs significantly better. As has been mentioned earlier, it is easy to come up with variations of the basic idea of move-to-front. Here are a few. 1. Move-ahead-k. The element of A matched by the current symbol is moved ahead k positions instead of all the way to the front of A. The parameter k can be specified by the user, with a default value of either n or 1. This tends to reduce performance (i.e., to increase the average size of the elements of C) for inputs that satisfy the concentration property, but it works better for other inputs. Notice that assigning k = n is identical to move-to-front. The case k = 1 is especially simple, since it only requires swapping an element of A with the one preceding it.  Exercise 1.11: Use move-ahead-k to encode each of the strings abcddcbamnopponm and abcdmnopabcdmnop twice, with k = 1 and k = 2. 48 1. Basic Techniques 2. Wait-c-and-move. An element of A is moved to the front only after it has been matched c times to symbols from the input stream (not necessarily c consecutive times). Each element of A should have a counter associated with it, to count the number of matches. This method makes sense in implementations where moving and rearranging elements of A is slow. 3. Normally, a symbol read from the input is a byte. If the input stream consists of text, however, it may make sense to treat each word, not each character, as a symbol. Consider the simple case where the input consists of just lowercase letters, spaces, and one end-of-text marker at the end. We can define a word as a string of letters followed by a space or by the end-of-text marker. The number of words in this case can be huge, so the alphabet list A should start empty, and words should be added as they are being input and encoded. We use the text theboyonmyrightistherightboy as an example. The first word input is the. It is not found in A, since A is empty, so it is added to A. The encoder emits 0 (the number of words preceding the in A) followed by the. The decoder also starts with an empty A. The 0 tells it to select the first word in A, but since A is empty, the decoder knows to expect the 0 to be followed by a word. It adds this word to A. The next word is boy. It is added to A, so A=(the, boy) and the encoder emits 1boy. The word boy is moved to the front of A, so A=(boy, the). The decoder reads the 1, which refers to the second word of A, but the decoder’s A has only one word in it so far. The decoder thus knows that a new word must follow the 1. It reads this word and adds it to the front of A. Table 1.16 summarizes the encoding steps for this example. List A may grow very large in this variant, but any practical implementation has to limit its size. This is why the last item of A (the least-recently-used item) has to be deleted when A exceeds its size limit. This is another difference between this variant and the basic move-to-front method.  Exercise 1.12: Decode theboyonmyrightistherightboy and summarize the steps in a table. Word A (before adding) A (after adding) Code emitted the () (the) 0the boy (the) (the, boy) 1boy on (boy, the) (boy, the, on) 2on my (on, boy, the) (on, boy, the, my) 3my right (my, on, boy, the) (my, on, boy, the, right) 4right is (right, my, on, boy, the) (right, my, on, boy, the, is) 5is the (is, right, my, on, boy, the) (is, right, my, on, boy, the) 5 right (the, is, right, my, on, boy) (the, is, right, my, on, boy) 2 boy (right, the, is, my, on, boy) (right, the, is, my, on, boy) 5 (boy, right, the, is, my, on) Table 1.16: Encoding Multiple-Letter Words. 1.6 Scalar Quantization 49 “I’ve been looking,” Lana Lee said, becoming a grave personnel manager, “for the right boy for this job for several days.” She put her hands in the pockets of her leather overcoat and looked into the sunglasses. —John Kennedy Toole, A Confederacy of Dunces 1.6 Scalar Quantization The dictionary definition of the term “quantization” is “to restrict a variable quantity to discrete values rather than to a continuous set of values.” In the field of data compression, quantization is employed in two ways: 1. If the data to be compressed is in the form of large numbers, quantization is used to convert it to small numbers. Small numbers take less space than large ones, so quantization generates compression. On the other hand, small numbers generally contain less information than large numbers, so quantization results in lossy compression. 2. If the data to be compressed is analog (i.e., a voltage that changes with time) quantization is used to digitize it into small numbers. The smaller the numbers the better the compression, but also the greater the loss of information. This aspect of quantization is exploited by several speech compression methods. In the discussion here we assume that the data to be compressed is in the form of numbers, and that it is input, number by number, from an input stream (or a source). Section 7.19 discusses a generalization of discrete quantization to cases where the data consists of sets (called vectors) of numbers rather than of individual numbers. The first example is naive discrete quantization of an input stream of 8-bit numbers. We can simply delete the least-significant four bits of each data item. This is one of those rare cases where the compression factor (=2) is known in advance and does not depend on the data. The input data consists of 256 different symbols, while the output data consists of just 16 different symbols. This method is simple but not very practical because too much information is lost in order to get the unimpressive compression factor of 2. In order to develop a better approach we assume again that the data consists of 8-bit numbers, and that they are unsigned. Thus, input symbols are in the range [0, 255] (if the input data is signed, input symbols have values in the range [-128, +127]). We select a spacing parameter s and compute the sequence of uniform quantized values 0, s, 2s,. . . ,ks, such that (k + 1)s > 255 and ks . 255. Each input symbol S is quantized by converting it to the nearest value in this sequence. Selecting s = 3, e.g., produces the uniform sequence 0,3,6,9,12,. . . ,252,255. Selecting s = 4 produces 0,4,8,12,. . . ,252,255 (since the next multiple of 4, after 252, is 256). A similar approach is to select quantized values such that any number in the range [0, 255] will be no more than d units distant from one of the data values that are being quantized. This is done by dividing the range into segments of size 2d+1 and centering them on the range [0, 255]. If, e.g., d = 16, then the range [0, 255] is divided into seven segments of size 33 each, with 25 numbers remaining. We can thus start the first segment 12 numbers from the start of the range, which produces the 10-number sequence 12, 33, 45, 78, 111, 144, 177, 210, 243, and 255. Any number in the range [0, 255] is at most 50 1. Basic Techniques 16 units distant from any of these numbers. If we want to limit the quantized sequence to just eight numbers (so each can be expressed in 3 bits) we can apply this method to compute the sequence 8, 41, 74, 107, 140, 173, 206, and 239. The quantized sequences above make sense in cases where each symbol appears in the input data with equal probability (cases where the source is memoryless). If the input data is not uniformly distributed, the sequence of quantized values should be distributed in the same way as the data. bbbbbbbb bbbbbbbb 1 1 . 10 2 . 11 3 . 100 4 100|000 32 101 5 101|000 40 110 6 110|000 48 111 7 111|000 56 100|0 8 100|0000 64 101|0 10 101|0000 80 110|0 12 110|0000 96 111|0 14 111|0000 112 100|00 16 100|00000 128 101|00 20 101|00000 160 110|00 24 110|00000 192 111|00 28 111|00000 224 Table 1.17: A Logarithmic Quantization Table. Imagine, e.g., an input stream of 8-bit unsigned data items most of which are zero or close to zero and few are large. A good sequence of quantized values for such data should have the same distribution, i.e., many small values and few large ones. One way of computing such a sequence is to select a value for the length parameter l and to construct a “window” of the form 1 b . . . bb    l , where each b is a bit, and place it under each of the 8 bit positions of a data item. If the window sticks out on the right, some of the l bits are truncated. As the window is shifted to the left, bits of 0 are appended to it. Table 1.17 illustrates this construction with l = 2. It is easy to see how the resulting quantized values start with initial spacing of one unit, continue with spacing of two units and four units, until the last four values are spaced by 32 units. The numbers 0 and 255 should be manually added to such a quasi-logarithmic sequence to make it more general. Scalar quantization is an example of a lossy compression method, where it is easy to control the trade-off between compression ratio and the amount of loss. However, because it is so simple, its use is limited to cases where much loss can be tolerated. Many image compression methods are lossy, but scalar quantization is not suitable for image compression because it creates annoying artifacts in the decompressed image. Imagine 1.7 Recursive Range Reduction 51 an image with an almost uniform area where all pixels have values 127 or 128. If 127 is quantized to 111 and 128 is quantized to 144, then the result, after decompression, may resemble a checkerboard where adjacent pixels alternate between 111 and 144. This is why practical algorithms use vector quantization, instead of scalar quantization, for lossy (and sometimes lossless) compression of images and sound. See also Section 7.3. 1.7 Recursive Range Reduction In their original 1977 paper [Ziv and Lempel 77], Lempel and Ziv have proved that their dictionary method can compress data to the entropy, but they pointed out that vast quantities of data would be needed to approach ideal compression. Other algorithms, most notably PPM (Section 5.14), suffer from the same problem. Often, a user is willing to sacrifice compression performance in favor of easy implementation and the knowledge that the performance of the algorithm is independent of the quantity of the data. The recursive range reduction (3R) method described here generates decent compression, is easy to implement, and its performance is independent of the amount of data to be compressed. These features make it an attractive candidate for compression in embedded systems, low-cost microcontrollers, and other applications where space is limited or resources are constrained. The method is the brainchild of Yann Guidon who described it in [3R 06]. The method is distantly related to variable-length codes. The method is described first for a sorted list of integers, where its application is simplest and no recursion is required. We term this version “range reduction” (RR). We then show how RR can be extended to a recursive version (RRR or 3R) that can be applied to any set of integers. Given a list of integers, we first eliminate the effects of the sign bit. We either rotate each integer to move the sign bit to the least-significant position (a folding) or add an offset to all the integers, so they become nonnegative. The list is then sorted in nonincreasing order. The first element of the list is the largest one, and we assume that its most-significant bit is a 1 (i.e., it has no extra zeros on the left). It is obvious that the integers that follow the first element may have some mostsignificant zero bits, and the heart of the RR method is to eliminate most of those bits while leaving enough information for the decoder to restore them. The first item in the compressed stream is a header with the length L of the largest integer. This length is stored in a fixed-size field whose size is sufficient for any data that may be encountered. In the examples here, this size is four bits, allowing for integers up to 16 bits long. Once the decoder inputs L, it knows the length of the first integer. The MSB of this integer is a 1, so this bit can be eliminated and the first integer can be emitted in L - 1 bits. The next integer is written on the output as an L-bit number, thereby allowing the decoder to read it unambiguously. The decoder then checks the most-significant bits of the second integer. If the leftmost k bits are zeros, then the decoder knows that the next integer (i.e., the third one) was written without its k leftmost zeros and thus occupies the next L - k bits in the compressed stream. This is how RR compresses a sorted list of nonnegative integers. The compression efficiency depends on the data, but not on the amount of data available for compression. The method does not improve by applying it to vast quantities of data. 52 1. Basic Techniques Table 1.18 is an example. The data consists of ten 7-bit integers. The 4-bit length field is set to 6, indicating 7-bit integers (notice that L cannot be zero, so 6 indicates a length of seven bits). The first integer is emitted minus its leftmost bit and the next integer is output in its entirety. The third integer is also emitted as is, but its leftmost zero (listed in bold) indicates to the decoder that the following (fourth) integer will have only six bits. Data RR code Rice code 0110 = 6 1011101 011101 000001 1101 1001011 1001011 00001 1011 0110001 0110001 0001 0001 0101100 101100 001 1100 0001110 001110 1 1110 0001101 1101 1 1101 0001100 1100 1 1100 0001001 1001 1 1001 0000010 0010 1 0010 0000001 01 1 0001 70 53 64 bits Table 1.18: Example of Range Reduction. The total size of the ten integers is 70 bits, and this should be compared with the 53 bits created by RR and the 64 bits resulting from the Rice codes (with n = 4) of the same integers (see Section 10.9 for Rice codes). Compression is not impressive, but it is obvious that it is not affected by the amount of data. The limited experience available with RR seems to indicate (but a rigorous proof is still needed) that this method performs best on data that decreases exponentially, where each integer is about half the size of its predecessor. In such a list, each number is one bit shorter than its predecessor. Worst results should be obtained with data that either decreases slowly, such as (900, 899, 898, 897, 896), or that decreases fast, such as (100000, 1000, 10, 1). These cases are illustrated in Table 1.19. Notice that the header is the only side information sent to the decoder and that RR never expands the data. Now for unsorted data. Given an unsorted list of integers, we can sort it, compress it with RR, and then include side information about the sort, so that the decoder can unsort the data correctly after decompressing it. An efficient method is required to specify the relation between a list of numbers and its sorted version, and the method described here is referred to as recursive range reduction or 3R. The 3R algorithm involves the creation of a binary tree where each path from the root to a leaf is a nonincreasing list of integers. Each path is encoded with 3R separately, which requires a recursive algorithm (hence the word “recursive” in the algorithm’s name). Figure 1.20 shows an example. Given the five unsorted integers A through E, a binary tree is constructed where each pair of consecutive integers becomes a subtree at the lowest level. Each of the five paths from the root to a leaf is a sorted list, and 1.7 Recursive Range Reduction 53 Data RR code Data RR code 0110 = 6 1101 = 13 1000010 000010 10011100010000 0011100010000 1000001 1000001 00001111101000 00001111101000 1000000 1000000 00000001100100 0001100100 0111111 0111111 00000000001010 0001010 0111110 111110 00000000000001 0001 0111101 111101 0111100 111101 49 49 70 52 Table 1.19: Range Reduction Worst Performance. it is compressed with 3R. It is obvious that writing the compressed results of all five lists on the output would normally cause expansion, so only certain nodes are actually written. The rule is to follow every path from the root and select the nodes along edges that go to the left (or only those that go to the right). Thus, in our example, we write the following values on the output: (1) a header with the length of the root, (2) the 3R-encoded path (root, A+B, A), (3) the value of node C [after the path (root, C+D+E, C) is encoded], and (4) the value of node D [after the path (root, C+D+E, D+E, D) is encoded]. A+B+C+D+E A+B C+D+E A B C D+E D E Figure 1.20: A Binary Tree Encoded With 3R. The decoder reads the header and first path (root, A+B, A). It decodes the path to obtain the three values and subtracts (A + B) - A to obtain B. It then subtracts Root-(A+B) to obtain C+D+E. It inputs C and now it can decode the path (root, C+D+E, C). Next, the decoder subtracts C from C+D+E to obtain D+E, it inputs D, and decodes path (root, C+D+E, D+E, D). One more subtraction yields E. This sounds like much work for very little compression, and it is! The 3R method is not the most efficient compression algorithm, but it may find its applications. The nice feature of the tree structure described here is that the number of nodes written on the output equals the size of the original data. In our example there are five data items and five 3R codes written on the output. It’s easy to see why this is generally true. Given a list of n data items, we observe that they end up as the leaves of the binary tree. What is eventually written on the output is half the nodes at each level of the tree. A complete binary tree with n leaves has 2n nodes, so half of this number, n nodes, ends up on the output. 54 1. Basic Techniques Cyan Magenta Yellow Black Figure 1.21: A Histogram. It seems that 3R would be suitable for applications where the data does not fit any of the standard statistical distributions and yet is not random. One example of such data is the histogram of an image (Figure 1.21 shows the CMYK histograms of the Lena image, Figure 7.55). On the other hand, 3R is restricted to nonnegative integers and the encoder requires two passes, one for constructing the tree and one for traversing it and collecting nodes. It has already been mentioned that RR never expands the data. However, when fed with random data, 3R generates output that grows by almost one bit per data item. Because of this feature, its developer refers to 3R as a nearly entropy encoder, not a real entropy encoder. Explanation. Because of the summations, random data are “averaged” so after a few tree levels, this is like almost-monotonous data. These data are then further summed together, still adding one bit of size per level but halving the number of sums each time, so in the end it is equivalent to one bit per input element. —Yann Guidon (private communication) Compression algorithms are often described as squeezing, squashing, crunching or imploding data, but these are not very good descriptions of what is actually happening. —James D. Murray and William Vanryper (1994) 2 Basic VL Codes The dates of most of the important historical events are known, but not always very precisely. We know that Kublai Khan, grandson of Ghengis Khan, founded the Yuan dynasty in 1280 (it lasted until 1368), but we don’t know precisely (i.e., the month, day and hour) when this act took place. A notable exception to this state of affairs is the modern age of telecommunications, a historical era whose birth is known precisely, up to the minute. On Friday, 24 May 1844, at precisely 9:45 in the morning, Samuel Morse inaugurated the age of modern telecommunications by sending the first telegraphic message in his new code. The message was sent over an experimental line funded by the American Congress from the Supreme Court chamber in Washington, DC to the B & O railroad depot in Baltimore, Maryland. Taken from the Bible (Numbers 23:23), the message was “What hath God wrought?” It had been suggested to Morse by Annie Ellsworth, the young daughter of a friend. It was prerecorded on a paper tape, was sent to a colleague in Baltimore, and was then decoded and sent back by him to Washington. An image of the paper tape can be viewed at [morse-tape 06]. Morse was born near Boston and was educated at Yale. We would expect the inventor of the telegraph (and of such a sophisticated code) to have been a child prodigy who tinkered with electricity and gadgets from an early age (the electric battery was invented when Morse was nine years old). Instead, Morse became a successful portrait painter with more than 300 paintings to his credit. It wasn’t until 1837 that the 46-year-old Morse suddenly quit his painting career and started thinking about communications and tinkering with electric equipment. It is not clear why he made such a drastic career change at such an age, but it is known that two large, wall-size paintings that he made for the Capitol building in DOI 10.1007/978-1-84882-903-9_2, © Springer-Verlag London Limited 2010 D. Salomon, G. Motta, Handbook of Data Compression, 5th ed. 56 2. Basic VL Codes Washington, DC were ignored by museum visitors and rejected by congressmen. It may have been this disappointment that gave us the telegraph and the Morse code. Given this background, it is easy to imagine how the 53-year-old Samuel Morse felt on that fateful day, Friday, 24 May 1844, as he sat hunched over his mysterious apparatus, surrounded by a curious crowd of onlookers, some of whom had only a vague idea of what he was trying to demonstrate. He must have been very anxious, because his telegraph project, his career, and his entire future depended on the success of this one test. The year before, the American Congress awarded him $30,000 to prepare this historical test and prove the value of the electric telegraph (and thus also confirm the ingenuity of yankees), and here he is now, dependent on the vagaries of his batteries, on the new, untested 41-mile-long telegraph line, and on a colleague in Baltimore. Fortunately, all went well. The friend in Baltimore received the message, decoded it, and resent it within a few minutes, to the great relief of Morse and to the amazement of the many congressmen assembled around him. The Morse code, with its quick dots and dashes (Table 2.1), was extensively used for many years, first for telegraphy, and beginning in the 1890s, for early radio communications. The development of more advanced communications technologies in the 20th century displaced the Morse code, which is now largely obsolete. Today, it is used for emergencies, for navigational radio beacons, land mobile transmitter identification, and by continuous wave amateur radio operators. A .- N -. 1 .---- Period .-.-.- B -... O --- 2 ..--- Comma --..-- C -.-. P .--. 3 ...-- Colon ---... Ch ---- Q --.- 4 ....- Question mark ..--.. D -.. R .-. 5 ..... Apostrophe .----. E . S ... 6 -.... Hyphen -....- F ..-. T - 7 --... Dash -..-. G --. U ..- 8 ---.. Parentheses -.--.- H .... V ...- 9 ----. Quotation marks .-..-. I .. W .-- 0 ----- J .--- X -..- K -.- Y -.-- L .-.. Z --.. M -- Table 2.1: The Morse Code for English. Our interest in the Morse code is primarily with a little-known aspect of this code. In addition to its advantages for telecommunications, the Morse code is also an early example of text compression. The various dot-dash codes developed by Morse (and possibly also by his associate, Alfred Vail) have different lengths, and Morse intuitively assigned the short codes (a single dot and a single dash) to the letters E and T, the longer, four dots-dashes, he assigned to Q, X, Y, and Z. The even longer, five dotsdashes codes, were assigned to the 10 digits, and the longest codes (six dots and dashes) 57 became those of the punctuation marks. Morse also specified that the signal for error is eight consecutive dots, in response to which the receiving operator should delete the last word received. It is interesting to note that Morse was not the first to think of compression (in terms of time saving) by means of a code. The well-known Braille code for the blind was developed by Louis Braille in the 1820s and is still in common use today. It consists of groups (or cells) of 3 Χ 2 dots each, embossed on thick paper. Each of the six dots in a group may be flat or raised, implying that the information content of a group is equivalent to six bits, resulting in 64 possible groups. The letters, digits, and common punctuation marks do not require all 64 codes, which is why the remaining groups may be used to code common words—such as and, for, and of—and common strings of letters—such as ound, ation, and th. The Morse code has another feature that makes it relevant to us. Because the individual codes have different lengths, there must be a way to identify the end of a code. Morse solved this problem by requiring accurate relative timing. If the duration of a dot is taken to be one unit, then that of a dash is three units, the space between the dots and dashes of one character is one unit, the space between characters is three units, and the interword space is six units (five for automatic transmission). This chapter and the two following it are concerned with the use of variable-length codes to compress digital data. With these codes, it is important not to have any extra spaces. In fact, there is no such thing as a space, because computers use only zeros and 1’s. Thus, when a string of data symbols is compressed by assigning codewords to the symbols, the codewords (whose lengths vary) are concatenated into a long binary string without any spaces or separators. Such variable-length codes must therefore be designed to allow for unambiguous reading. Somehow, the decoder should be able to read bits and identify the end of each codeword. Such codes are referred to as uniquely decodable or uniquely decipherable (UD). Variable-length codes have become important in many areas of computer science. This chapter and the two following it are related and constitute a survey of this important topic. They present the principles underlying this type of codes and describe the important classes of variable-length codes. Many examples illustrate the applications of these codes to data compression. These three related chapters are devoted to the codes, which is why they describe the codes rather than actual compression algorithms. The remainder of the book is devoted to many actual methods, algorithms, and techniques for compressing data. The term representation is central to our discussion. A number can be represented in decimal, binary, or any other number base (or number system, see Section 3.19). Mathematically, a representation is a bijection (or a bijective function) of an infinite, countable set S1 of strings onto another set S2 of strings (in practice, S2 consists of binary strings, but it may also be ternary or based on other number systems), such that any concatenation of any elements of S2 is UD. The elements of S1 are called data symbols and those of S2 are codewords. Set S1 is an alphabet and set S2 is a code. An interesting example is the standard binary notation. We normally refer to it as the binary representation of the integers, but according to the definition above it is not a representation because it is not UD. It is easy to see, for example, that a string of binary codewords that starts with 11 can be either two consecutive 1’s or the code of 3. 58 2. Basic VL Codes A function f : X . Y is said to be bijective, if for every y . Y , there is exactly one x . X such that f(x) = y. Figure 4.19 and Table 4.22 list several variable-length UD codes assigned to the 26 letters of the English alphabet. This group of three related chapters starts with several introductory sections (Sections 2.1 through 2.6) that discuss information theory concepts such as entropy and redundancy, and concepts that are used throughout the text, such as prefix codes, complete codes, and universal codes. The remainder of the group deals mostly with block-to-variable codes, although its first part deals with the Tunstall codes and other variable-to-block codes. It concentrates on the codes themselves, not on the compression algorithms. Thus, the individual sections describe various variable-length codes and classify them according to their structure and organization. The main techniques employed to design variable-length codes are the following: The phased-in codes (Section 2.9) are a minor extension of fixed-length codes and may contribute a little to the compression of a set of consecutive integers by changing the representation of the integers from fixed n bits to either n or n - 1 bits (recursive phased-in codes are also described). Self-delimiting codes. These are intuitive variable-length codes—mostly due to Gregory Chaitin, the originator of algorithmic information theory—where a code signals its end by means of extra flag bits. The self-delimiting codes of Section 2.12 are inefficient and are not used in practice. Prefix codes. Such codes can be read unambiguously (they are uniquely decodable, or UD codes) from a long string of codewords because they have a special property (the prefix property) which is stated as follows: Once a bit pattern is assigned as the code of a symbol, no other codes can start with that pattern. The most common example of prefix codes are the Huffman codes (Section 5.2). Other important examples are the unary, start-step-stop, and start/stop codes (Sections 3.2 and 3.3, respectively). Codes that include their own length. One way to construct a UD code for the integers is to start with the standard binary representation of an integer and prepend to it its length L1. The length may also have variable length, so it has to be encoded in some way or have its length L2 prepended. The length of an integer n equals approximately log n (where the logarithm base is the same as the number base of n), which is why such methods are often called logarithmic ramp representations of the integers. The most common examples of this type of codes are the Elias codes (Section 3.4), but other types are also presented. They include the Levenstein code (Section 3.6), Even– Rodeh code (Section 3.7), punctured Elias codes (Section 3.8), the ternary comma code (Section 3.10), Stout codes (Section 3.12), Boldi–Vigna (zeta) codes (Section 3.13), and Yamamoto’s recursive code (Section 3.14). Suffix codes (codes that end with a special flag). Such codes limit the propagation of an error and are therefore robust. An error in a codeword affects at most that codeword and the one or two codewords following it. Most other variable-length codes 59 sacrifice data integrity to achieve short codes, and are fragile because a single error can propagate indefinitely through a sequence of concatenated codewords. The taboo codes of Section 3.16 are UD because they reserve a special string (the taboo) to indicate the end of the code. Wang’s flag code (Section 3.17) is also included in this category. Note. The term “suffix code” is ambiguous. It may refer to codes that end with a special bit pattern, but it also refers to codes where no codeword is the suffix of another codeword (the opposite of prefix codes). The latter meaning is used in Section 4.5, in connection with bidirectional codes. Flag codes. A true flag code differs from the suffix codes in one interesting aspect. Such a code may include the flag inside the code, as well as at its right end. The only example of a flag code is Yamamoto’s code, Section 3.18. Codes based on special number bases or special number sequences. We normally use decimal numbers, and computers use binary numbers, but any integer greater than 1 can serve as the basis of a number system and so can noninteger (real) numbers. It is also possible to construct a sequence of numbers (real or integer) that act as weights of a numbering system. The most important examples of this type of variable-length codes are the Fibonacci (Section 3.20), Goldbach (Section 3.22), and additive codes (Section 3.23). The Golomb codes of Section 3.24 are designed in a special way. An integer parameter m is selected and is used to encode an arbitrary integer n in two steps. In the first step, two integers q and r (for quotient and remainder) are computed from n such that n can be fully reconstructed from them. In the second step, q is encoded in unary and is followed by the binary representation of r, whose length is implied by the parameter m. The Rice code of Section 3.25 is a special case of the Golomb codes where m is an integer power of 2. The subexponential code (Section 3.26) is related to the Rice codes. Codes ending with “1” are the topic of Section 3.27. In such a code, all the codewords end with a 1, a feature that makes them the natural choice in special applications. Variable-length codes are designed for data compression, which is why implementors select the shortest possible codes. Sometimes, however, data reliability is a concern, and longer codes may help detect and isolate errors. Thus, Chapter 4 discusses robust codes. Section 4.3 presents synchronous prefix codes. These codes are useful in applications where it is important to limit the propagation of errors. Bidirectional (or reversible) codes (Sections 4.5 and 4.6) are also designed for increased reliability by allowing the decoder to read and identify codewords either from left to right or in reverse. The discussion in this chapter starts with codes, prefix codes, and information theory concepts. This is followed by a description of basic codes such as variable-toblock codes, phased-in codes, and the celebrated Huffman code. 60 2. Basic VL Codes 2.1 Codes, Fixed- and Variable-Length A code is a symbol that stands for another symbol. At first, this idea seems pointless. Given a symbol S, what is the use of replacing it with another symbol Y ? However, it is easy to find many important examples of the use of codes. Here are a few. Any language and any system of writing are codes. They provide us with symbols Y that we use in order to express our thoughts S. Acronyms and abbreviations can be considered codes. Thus, the string IBM is a symbol that stands for the much longer symbol “International Business Machines” and the well-known French university ΄Ecole Sup΄erieure D’΄electricit΄e is known to many simply as Sup΄elec. Cryptography is the art and science of obfuscating messages. Before the age of computers, a message was typically a string of letters and was encrypted by replacing each letter with another letter or with a number. In the computer age, a message is a binary string (a bitstring) in a computer, and it is encrypted by replacing it with another bitstring, normally of the same length. Error control. Messages, even secret ones, are often transmitted over communications channels and may become damaged, corrupted, or garbled on their way from transmitter to receiver. We often experience low-quality, garbled telephone conversations. Even experienced pharmacists often find it difficult to read and understand a handwritten prescription. Computer data stored on magnetic disks may become corrupted because of exposure to magnetic fields or extreme temperatures. Music and movies recorded on optical discs (CDs and DVDs) may become unreadable because of scratches. In all these cases, it helps to augment the original data with error-control codes. Such codes—formally titled channel codes, but informally known as error-detecting or errorcorrecting codes—employ redundancy to detect and even correct, certain types of errors. ASCII and Unicode. These are character codes that make it possible to store characters of text as bitstrings in a computer. The ASCII code, which dates back to the 1960s [ascii-wiki 09], assigns 7-bit codes to 128 characters including 26 letters (upper- and lowercase), the 10 digits, certain punctuation marks, and several control characters. The Unicode project assigns 16-bit codes to many characters, and has a provision for even longer codes. The long codes make it possible to store and manipulate many thousands of characters, taken from many languages and alphabets (such as Greek, Cyrillic, Hebrew, Arabic, and Indic), and including punctuation marks, diacritics, mathematical symbols, technical symbols, arrows, and dingbats. The last example illustrates the use of codes in the field of computers and computations. Mathematically, a code is a mapping. Given an alphabet of symbols, a code maps individual symbols or strings of symbols to codewords, where a codeword is a string of bits, a bitstring. The process of mapping a symbol to a codeword is termed encoding and the reverse process is known as decoding. Codes can have a fixed or variable length, and can be static or adaptive (dynamic). A static code is constructed once and never changes. ASCII and Unicode are examples of such codes. A static code can also have variable length, where short codewords are 2.1 Codes, Fixed- and Variable-Length 61 assigned to the commonly-occurring symbols. A variable-length, static code is normally designed based on the probabilities of the individual symbols. Each type of data has different probabilities and may benefit from a different code. The Huffman method (Section 5.2) is an example of an excellent variable-length, static code that can be constructed once the probabilities of all the symbols in the alphabet are known. In general, static codes that are also variable-length can match well the lengths of individual codewords to the probabilities of the symbols. Notice that the code table must normally be included in the compressed file, because the decoder does not know the symbols’ probabilities (the model of the data) and so has no way to construct the codewords independently. A dynamic code varies over time, as more and more data is read and processed and more is known about the probabilities of the individual symbols. The dynamic (adaptive) Huffman algorithm [Section 5.3] is a method that employs such a code. Fixed-length codes are known as block codes. They are easy to implement in software. It is easy to replace an original symbol with a fixed-length code, and it is equally easy to start with a string of such codes and break it up into individual codes that are then replaced by the original symbols. There are cases where variable-length codes (VLCs) have obvious advantages. As their name implies, VLCs are codes that have different lengths. They are also known as variable-length codes. A set of such codes consists of short and long codewords. The following is a short list of important applications where such codes are commonly used. Data compression (or source coding). Given an alphabet of symbols where certain symbols occur often in messages, while other symbols are rare, it is possible to compress messages by assigning short codes to the common symbols and long codes to the rare symbols. This is an important application of variable-length codes. The Morse code for telegraphy, originated in the 1830s by Samuel Morse and Alfred Vail, exploits the same idea. It assigns short codes to commonly-occurring letters (the code of E is a dot and the code of T is a dash) and long codes to rare letters and punctuation marks (--.- to Q, --.. to Z, and --..-- to the comma). Processor design. Part of the architecture of any computer is an instruction set and a processor that fetches instructions from memory and executes them. It is easy to handle fixed-length instructions, but modern computers normally have instructions of different sizes. It is possible to reduce the overall size of programs by designing the instruction set such that commonly-used instructions are short. This also reduces the processor’s power consumption and physical size and is especially important in embedded processors, such as processors designed for digital signal processing (DSP). Country calling codes. ITU-T recommendation E.164 is an international standard that assigns variable-length calling codes to many countries such that countries with many telephones are assigned short codes and countries with fewer telephones are assigned long codes. These codes also obey the prefix property (Section 2.2) which means that once a calling code C has been assigned, no other calling code will start with C. The International Standard Book Number (ISBN) is a unique number assigned to a book, to simplify inventory tracking by publishers and bookstores. The ISBN numbers are assigned according to an international standard known as ISO 2108 (1970). One 62 2. Basic VL Codes component of an ISBN is a country code, that can be between one and five digits long. This code also obeys the prefix property. Once C has been assigned as a country code, no other country code will start with C. VCR Plus+ (also known as G-Code, VideoPlus+, and ShowView) is a prefix, variable-length code for programming video recorders. A unique number, a VCR Plus+, is computed for each television program by a proprietary algorithm from the date, time, and channel of the program. The number is published in television listings in newspapers and on the Internet. To record a program on a VCR, the number is located in a newspaper and is typed into the video recorder. This programs the recorder to record the correct channel at the right time. This system was developed by Gemstar-TV Guide International [Gemstar 06]. I gave up on new poetry myself thirty years ago, when most of it began to read like coded messages passing between lonely aliens on a hostile world. —Russell Baker 2.2 Prefix Codes Encoding a string of symbols ai with VLCs is easy. No clever methods or algorithms are needed. The software reads the original symbols ai one by one and replaces each ai with its binary, variable-length code ci. The codes are concatenated to form one (normally long) bitstring. The encoder either includes a table with all the pairs (ai, ci) or it executes a procedure to compute code ci from the bits of symbol ai. Decoding is slightly more complex, because of the different lengths of the codes. When the decoder reads the individual bits of VLCs from a bitstring, it has to know either how long each code is or where each code ends. This is why a set of variable-length codes has to be carefully selected and why the decoder has to be taught about the codes. The decoder either has to have a table of all the valid codes, or it has to be told how to identify valid codes. We start with a simple example. Given the set of four codes a1 = 0, a2 = 01, a3 = 011, and a4 = 111 we easily encode the message a2a3a3a1a2a4 as the bitstring 01|011|011|0|01|111. This string can be decoded unambiguously, but not easily. When the decoder inputs a 0, it knows that the next symbol is either a1, a2, or a3, but the decoder has to input more bits to find out how many 1’s follow the 0 before it can identify the next symbol. Similarly, given the bitstring 011 . . . 111, the decoder has to read the entire string and count the number of consecutive 1’s before it finds out how many 1’s (zero, one, or two 1’s) follow the single 0 at the beginning. We say that such codes are not instantaneous. In contrast, the following set of VLCs a1 = 0, a2 = 10, a3 = 110, and a4 = 111 is similar and is also instantaneous. Given a bitstring that consists of these codes, the decoder reads consecutive 1’s until it has read three 1’s (an a4) or until it has read another 0. Depending on how many 1’s precede the 0 (zero, one, or two 1’s), the decoder knows whether the next symbol is a1, a2, or a3. The 0 acts as a separator, which is why instantaneous codes are also known as comma codes. The rules that drive the decoder can be considered a finite automaton or a decision tree. 2.3 VLCs, Entropy, and Redundancy 63 The next example is similar. We examine the set of VLCs a1 = 0, a2 = 10, a3 = 101, and a4 = 111. Only the code of a3 is different, but a little experimenting shows that this set of VLCs is bad because it is not uniquely decodable (UD). Given the bitstring 0101111. . . , it can be decoded either as a1a3a4 . . . or a1a2a4 . . .. This observation is crucial because it points the way to the construction of large sets of VLCs. The set of codes above is bad because 10, the code of a2, is also the prefix of the code of a3. When the decoder reads 10. . . , it often cannot tell whether this is the code of a2 or the start of the code of a3. Thus, a useful, practical set of VLCs has to be instantaneous and has to satisfy the following prefix property. Once a code c is assigned to a symbol, no other code should start with the bit pattern c. Prefix codes are also referred to as prefix-free codes, prefix condition codes, or instantaneous codes. The following results can be proved: (1) A code is instantaneous if and only if it is a prefix code. (2) The set of UD codes is larger than the set of instantaneous codes (i.e., there are UD codes that are not instantaneous). (3) There is an instantaneous variable-length code with codeword lengths Li if and only if there is a UD code with these codeword lengths. The last of these results indicates that we cannot reduce the average word length of a variable-length code by using a UD code rather than an instantaneous code. Thus, there is no loss of compression performance if we restrict our selection of codes to instantaneous codes. A UD code that consists of r codewords of lengths li must satisfy the Kraft inequality (Section 2.5), but this inequality does not require a prefix code. Thus, if a code satisfies the Kraft inequality it is UD, but if it is also a prefix code, then it is instantaneous. This feature of a UD code being also instantaneous, comes for free, because there is no need to add bits to the code and make it longer. A prefix code (a set of codewords that satisfy the prefix property) is UD. Such a code is also complete if adding any codeword to it turns it into a non-UD code. A complete code is the largest UD code, but it also has a downside; it is less robust. If even a single bit is accidentally modified or deleted (or if a bit is somehow added) during storage or transmission, the decoder will lose synchronization and the rest of the transmission will be decoded incorrectly (see the discussion of robust codes in Chapter 4). While discussing UD and non-UD codes, it is interesting to note that the Morse code is non-UD (because, for example, the code of I is “..” and the code of H is “....”), so Morse had to make it UD by requiring accurate relative timing. 2.3 VLCs, Entropy, and Redundancy Understanding data compression and its codes must start with understanding information, because the former is based on the latter. Hence this short section that introduces a few important concepts from information theory (see the more detailed discussion in the Appendix). 64 2. Basic VL Codes Information theory is the creation, in the late 1940s, of Claude Shannon. Shannon tried to develop means for measuring the amount of information stored in a symbol without considering the meaning of the information. He discovered the connection between the logarithm function and information, and showed that the information content (in bits) of a symbol with probability p is -log2 p. If the base of the logarithm is e, then the information is measured in units called nats. If the base is 3, the information units are trits, and if the base is 10, the units are referred to as Hartleys. Information theory is concerned with the transmission of information from a sender (termed a source), through a communications channel, to a receiver. The sender and receiver can be persons or machines and the receiver may, in turn, act as a sender and send the information it has received to another receiver. The information is sent in units called symbols (normally bits, but in verbal communications the symbols are spoken words) and the set of all possible data symbols is an alphabet. The most important single factor affecting communications is noise in the communications channel. In verbal communications, this noise is literally noise. When trying to talk in a noisy environment, we may lose part of the discussion. In electronic communications, the channel noise is caused by imperfect hardware and by factors such as lightning, voltage fluctuations—old, high-resistance wires—sudden surge in temperature, and interference from machines that generate strong electromagnetic fields. The presence of noise implies that special codes should be used to increase the reliability of transmitted information. This is referred to as channel coding or, in everyday language, error-control codes. The second most important factor affecting communications is sheer volume. Any communications channel has a limited capacity. It can transmit only a limited number of symbols per time unit. An obvious way to increase the amount of data transmitted is to compress it before it is sent (in the source). Methods to compress data are therefore known as source coding or, in everyday language, data compression. The feature that makes it possible to compress data is the fact that individual symbols appear in our data files with different probabilities. One important principle (although not the only one) used to compress data is to assign variable-length codes to the individual data symbols such that short codes are assigned to the common symbols. Two concepts from information theory, namely entropy and redundancy, are needed in order to fully understand the application of VLCs to data compression. Roughly speaking, the term “entropy” as defined by Shannon is proportional to the minimum number of yes/no questions needed to reach the answer to some question. Another way of looking at entropy is as a quantity that describes how much information is included in a signal or event. Given a discrete random variable X that can have n values xi with probabilities Pi, the entropy H(X) of X is defined as H(X) = - n i=1 Pi log2 Pi. 2.3 VLCs, Entropy, and Redundancy 65 A detailed discussion of information theory is outside the scope of this book. Interested readers are referred to the many texts on this subject. Here, we will only show intuitively why the logarithm function plays such an important part in measuring information. Imagine a source that emits symbols ai with probabilities pi. We assume that the source is memoryless, i.e., the probability of a symbol being emitted does not depend on what has been emitted so far. We want to define a function I(ai) that will measure the amount of information gained when we discover that the source has emitted symbol ai. Function I will also measure our uncertainty as to whether the next symbol will be ai. Alternatively, I(ai) corresponds to our surprise in finding that the next symbol is ai. Clearly, our surprise at seeing ai emitted is inversely proportional to the probability pi (we are surprised when a low-probability symbol is emitted, but not when we notice a high-probability symbol). Thus, it makes sense to require that function I satisfies the following conditions: 1. I(ai) is a decreasing function of pi, and returns 0 when the probability of a symbol is 1. This reflects our feeling that high-probability events convey less information. 2. I(aiaj) = I(ai) + I(aj ). This is a result of the source being memoryless and the probabilities being independent. Discovering that ai was immediately followed by aj , provided us with the same information as knowing that ai and aj were emitted independently. Even those with a minimal mathematical background will immediately see that the logarithm function satisfies the two conditions above. This is the first example of the relation between the logarithm function and the quantitative measure of information. The next few paragraphs illustrate other connections between the two. Consider the case of person A selecting at random an integer N between 1 and 64 and person B having to guess N. What is the minimum number of yes/no questions that are needed for B to guess N? Those familiar with the technique of binary search know how to guess efficiently. Using this technique, B should divide the interval 1–64 in two, and should start by asking “is N between 1 and 32?” If the answer is no, then N is in the interval 33 to 64. This interval is then divided by two and B’s next question should be “is N between 33 and 48?” This process continues until the interval selected by B shrinks to a single number. It does not take much to see that exactly six questions are necessary to determine N. This is because 6 is the number of times 64 can be divided in half. Mathematically, this is equivalent to writing 6 = log2 64, which is why the logarithm is the mathematical function that quantifies information. What we call reality arises in the last analysis from the posing of yes/no questions. All things physical are information-theoretic in origin, and this is a participatory universe. —John Wheeler Another approach to the same problem is to consider a nonnegative integer N and ask how many digits does it take to express it. The answer, of course, depends on N. The greater N, the more digits are needed. The first 100 nonnegative integers (0 through 99) can be expressed by two decimal digits. The first 1000 such integers can be expressed by three digits. Again it does not take long to see the connection. The number of digits required to represent N equals approximately logN. The base of the logarithm is the 66 2. Basic VL Codes same as the base of the digits. For decimal digits, use base 10; for binary digits (bits), use base 2. If we agree that the number of digits it takes to express N is proportional to the information content of N, then again the logarithm is the function that gives us a measure of the information. As an aside, the precise length, in bits, of the binary (ί) representation of a positive integer n is 1 + log2 n (2.1) or, alternatively, log2(n + 1). When n is represented in any other number base b, its length is given by the same formula, but with the logarithm in base b instead of 2. Here is another observation that illuminates the relation between the logarithm and information. A 10-bit string can have 210 = 1,024 values. We say that such a string may contain one of 1,024 messages, or that the length of the string is the logarithm of the number of possible messages the string can convey. The following example sheds more light on the concept of entropy and will prepare us for the definition of redundancy. Given a set of two symbols a1 and a2, with probabilities P1 and P2, respectively, we compute the entropy of the set for various values of the probabilities. Since P1+P2 = 1, the entropy of the set is -P1 log2 P1-(1-P1) log2(1-P1) and the results are summarized in Table 2.2. When P1 = P2, at least one bit is required to encode each symbol, reflecting the fact that the entropy is at its maximum, the redundancy is zero, and the data cannot be compressed. However, when the probabilities are very different, the minimum number of bits required per symbol drops significantly. We may not be able to conceive a compression method that expresses each symbol in just 0.08 bits, but we know that when P1 = 99%, such compression is theoretically possible. P1 P2 Entropy 0.99 0.01 0.08 0.90 0.10 0.47 0.80 0.20 0.72 0.70 0.30 0.88 0.60 0.40 0.97 0.50 0.50 1.00 Table 2.2: Probabilities and Entropies of Two Symbols. In general, the entropy of a set of n symbols depends on the individual probabilities Pi and is largest when all n probabilities are equal. Data representations often include redundancies and data can be compressed by reducing or eliminating these redundancies. When the entropy is at its maximum, the data has maximum information content and therefore cannot be further compressed. Thus, it makes sense to define redundancy as a quantity that goes down to zero as the entropy reaches its maximum. The fundamental problem of communication is that of reproducing at one point either exactly or approximately a message selected at another point. —Claude Shannon (1948) 2.3 VLCs, Entropy, and Redundancy 67 To understand the definition of redundancy, we start with an alphabet of symbols ai, where each symbol appears in the data with probability Pi. The data is compressed by replacing each symbol with an li-bit-long code. The average code length is the sum Pili and the entropy (the smallest number of bits required to represent the symbols) is [-Pi log2 Pi]. The redundancy R of the set of symbols is defined as the average code length minus the entropy. Thus, R =i Pili -i [-Pi log2 Pi]. (2.2) The redundancy is zero when the average code length equals the entropy, i.e., when the codes are the shortest and compression has reached its maximum. Given a set of symbols (an alphabet), we can assign binary codes to the individual symbols. It is easy to assign long codes to symbols, but most practical applications require the shortest possible codes. Consider the four symbols a1, a2, a3, and a4. If they appear in our data strings with equal probabilities (= 0.25), then the entropy of the data is -4(0.25 log2 0.25) = 2. Two is the smallest number of bits needed, on average, to represent each symbol in this case. We can simply assign our symbols the four 2-bit codes 00, 01, 10, and 11. Since the probabilities are equal, the redundancy is zero and the data cannot be compressed below two bits/symbol. Next, consider the case where the four symbols occur with different probabilities as shown in Table 2.3, where a1 appears in the data (on average) about half the time, a2 and a3 have equal probabilities, and a4 is rare. In this case, the data has entropy -(0.49 log2 0.49+0.25 log2 0.25+0.25 log2 0.25+0.01 log2 0.01) . -(-0.050-0.5-0.5- 0.066) = 1.57. The smallest number of bits needed, on average, to represent each symbol has dropped to 1.57. Symbol Prob. Code1 Code2 a1 .49 1 1 a2 .25 01 01 a3 .25 010 000 a4 .01 001 001 Table 2.3: Variable-Length Codes. If we again assign our symbols the four 2-bit codes 00, 01, 10, and 11, the redundancy would be R = -1.57 + log24 = 0.43. This suggests assigning variable-length codes to the symbols. Code1 of Table 2.3 is designed such that the most common symbol, a1, is assigned the shortest code. When long data strings are transmitted using Code1, the average size (the number of bits per symbol) is 1 Χ 0.49 + 2 Χ 0.25 + 3 Χ 0.25 + 3 Χ 0.01 = 1.77, which is very close to the minimum. The redundancy in this case is R = 1.77 - 1.57 = 0.2 bits per symbol. An interesting example is the 20-symbol string a1a3a2a1a3a3a4a2a1a1a2a2a1a1a3a1a1a2a3a1, where the four symbols occur with approximately the right frequencies. Encoding this string with Code1 yields the 37 bits: 68 2. Basic VL Codes 1|010|01|1|010|010|001|01|1|1|01|01|1|1|010|1|1|01|010|1 (without the vertical bars). Using 37 bits to encode 20 symbols yields an average size of 1.85 bits/symbol, not far from the calculated average size. (The reader should bear in mind that our examples are short. To obtain results close to the best that’s theoretically possible, an input stream with at least thousands of symbols is needed.) However, the conscientious reader may have noticed that Code1 is bad because it is not a prefix code. Code2, in contrast, is a prefix code and can be decoded uniquely. Notice how Code2 was constructed. Once the single bit 1 was assigned as the code of a1, no other codes could start with 1 (they all had to start with 0). Once 01 was assigned as the code of a2, no other codes could start with 01. This is why the codes of a3 and a4 had to start with 00. Naturally, they became 000 and 001. Designing variable-length codes for data compression must therefore take into account the following two principles: (1) assign short codes to the more frequent symbols and (2) obey the prefix property. Following these principles produces short, unambiguous codes, but not necessarily the best (i.e., shortest) codes. In addition to these principles, an algorithm is needed to generate a set of shortest codes (ones with the minimum average size). The only input to such an algorithm is the frequencies of occurrence (or alternatively the probabilities) of the symbols of the alphabet. The well-known Huffman algorithm (Section 5.2) is such a method. Given a set of symbols whose probabilities of occurrence are known, this algorithm constructs a set of shortest prefix codes for the symbols. Notice that such a set is normally not unique and there may be several sets of codes with the shortest length. The beauty of code is much more akin to the elegance, efficiency and clean lines of a spiderweb. It is not the chaotic glory of a waterfall, or the pristine simplicity of a flower. It is an aesthetic of structure, design and order. —Charles Gordon Notice that a UD code does not have to be a prefix code. It is possible, for example, to designate the string 111 as a separator (a comma) to separate individual codewords of different lengths, provided that no codeword contains 111. Other examples of a nonprefix, variable-length codes are the C3 code (page 146) and the generalized Fibonacci C2 code (page 149). 2.4 Universal Codes Mathematically, a code is a mapping. It maps source symbols into codewords. Mathematically, a source of messages is a pair (M,P) where M is a (possibly infinite) set of messages and P is a function that assigns a nonzero probability to each message. A message is mapped into a long bitstring whose length depends on the quality of the code and on the probabilities of the individual symbols. The best that can be done is to compress a message to its entropy H. A code is universal if it compresses messages to codewords whose average length is bounded by C1(H + C2) where C1 and C2 are constants greater than or equal to 1, i.e., an average length that is a constant multiple of the entropy plus another constant. A universal code with large constants isn’t very useful. A code with C1 = 1 is called asymptotically optimal. 2.5 The Kraft–McMillan Inequality 69 A Huffman code often performs better than a universal code, but it can be used only when the probabilities of the symbols are known. In contrast, a universal code can be used in cases where only the ranking of the symbols’ probabilities is known. If we know that symbol a5 has the highest probability and a8 has the next largest one, we can assign the shortest codeword to a5 and the next longer codeword to a8. Thus, universal coding amounts to ranking of the source symbols. After ranking, the symbol with index 1 has the largest probability, the symbol with index 2 has the next highest one, and so on. We can therefore ignore the actual symbols and concentrate on their new indexes. We can assign one codeword to index 1, another codeword to index 2, and so on, which is why variable-length codes are often designed to encode integers (Section 3.1). Such a set of variable-length codes can encode any number of integers with codewords that have increasing lengths. Notice also that a set of universal codes is fixed and so doesn’t have to be constructed for each set of source symbols, a feature that simplifies encoding and decoding. However, if we know the probabilities of the individual symbols (the probability distribution of the alphabet of symbols), it becomes possible to tailor the code to the probability, or conversely, to select a known code whose codewords fit the known probability distribution. In all cases, the code selected (the set of codewords) must be uniquely decodable (UD). A non-UD code is ambiguous and therefore useless. 2.5 The Kraft–McMillan Inequality The Kraft–McMillan inequality is concerned with the existence of a uniquely decodable (UD) code. It establishes the relation between such a code and the lengths Li of its codewords. One part of this inequality, due to [McMillan 56], states that given a UD variablelength code, with n codewords of lengths Li, the lengths must satisfy the relation n i=1 2-Li . 1. (2.3) The other part, due to [Kraft 49], states the opposite. Given a set of n positive integers (L1, L2, . . . , Ln) that satisfy Equation (2.3), there exists an instantaneous variable-length code such that the Li are the lengths of its individual codewords. Together, both parts say that there is an instantaneous variable-length code with codeword lengths Li if and only if there is a UD code with these codeword lengths. The two parts do not say that a variable-length code is instantaneous or UD if and only if the codeword lengths satisfy Equation (2.3). In fact, it is easy to check the three individual code lengths of the code (0, 01, 011) and verify that 2-1 + 2-2 + 2-3 = 7/8. This code satisfies the Kraft–McMillan inequality and yet it is not instantaneous, because it is not a prefix code. Similarly, the code (0, 01, 001) also satisfies Equation (2.3), but is not UD. A few more comments on this inequality are in order: If a set of lengths Li satisfies Equation (2.3), then there exist instantaneous and UD variable-length codes with these lengths. For example (0, 10, 110). 70 2. Basic VL Codes A UD code is not always instantaneous, but there exists an instantaneous code with the same codeword lengths. For example, code (0, 01, 11) is UD but not instantaneous, while code (0, 10, 11) is instantaneous and has the same lengths. The sum of Equation (2.3) corresponds to the part of the complete code tree that has been used for codeword selection. This is why the sum has to be less than or equal to 1. This intuitive explanation of the Kraft–McMillan relation is explained in the next paragraph. We can gain a deeper understanding of this useful and important inequality by constructing the following simple prefix code. Given five symbols ai, suppose that we decide to assign 0 as the code of a1. Now all the other codes have to start with 1. We therefore assign 10, 110, 1110, and 1111 as the codewords of the four remaining symbols. The lengths of the five codewords are 1, 2, 3, 4, and 4, and it is easy to see that the sum 2-1 + 2-2 + 2-3 + 2-4 + 2-4 = 1 2 + 1 4 + 1 8 + 2 16 = 1 satisfies the Kraft–McMillan inequality. We now consider the possibility of constructing a similar code with lengths 1, 2, 3, 3, and 4. The Kraft–McMillan inequality tells us that this is impossible, because the sum 2-1 + 2-2 + 2-3 + 2-3 + 2-4 = 1 2 + 1 4 + 2 8 + 1 16 is greater than 1, and this is easy to understand when we consider the code tree. Starting with a complete binary tree of height 4, such as the tree of Figure 4.17, it is obvious that once 0 was assigned as a codeword, we have “used” one half of the tree and all future codes would have to be selected from the other half of the tree. Once 10 was assigned, we were left with only 1/4 of the tree. Once 110 was assigned as a codeword, only 1/8 of the tree remained available for the selection of future codes. Once 1110 has been assigned, only 1/16 of the tree was left, and that was enough to select and assign code 1111. However, once we select and assign codes of lengths 1, 2, 3, and 3, we have exhausted the entire tree and there is nothing left to select the last (4-bit) code from. The Kraft–McMillan inequality can be related to the entropy by observing that the lengths Li can always be written as Li = -log2 Pi +Ei, where Ei is simply the amount by which Li is greater than the entropy (the extra length of code i). This implies that 2-Li = 2(log2 Pi-Ei) = 2log2 Pi/2Ei = Pi/2Ei . In the special case where all the extra lengths are the same (Ei = E), the Kraft–McMillan inequality says that 1 . n i=1 Pi/2E = n i=1 Pi 2E = 1 2E =. 2E . 1 =. E . 0. An unambiguous code has nonnegative extra length, meaning its length is greater than or equal to the length determined by its entropy. 2.5 The Kraft–McMillan Inequality 71 Here is a simple example of the use of this inequality. Consider the simple case of n equal-length binary codewords. The size of each codeword is Li = log2 n, and the Kraft–McMillan sum is n 1 2-Li = n 1 2-log2 n = n 1 1 n = 1. The inequality is satisfied, so such a code is UD. A more interesting example is the case of n symbols where the first one is compressed and the second one is expanded. We set L1 = log2 n - a, L2 = log2 n + e, and L3 = L4 = · · · = Ln = log2 n, where a and e are positive. We show that e > a, which means that compressing a symbol by a factor a requires expanding another symbol by a larger factor. We can benefit from this only if the probability of the compressed symbol is greater than that of the expanded symbol. n 1 2-Li = 2-L1 + 2-L2 + n 3 2-log2 n = 2-log2 n+a + 2-log2 n-e + n 1 2-log2 n - 2 Χ 2-log2 n = 2a n + 2-e n + 1 - 2 n . The Kraft–McMillan inequality requires that 2a n + 2-e n + 1 - 2 n . 1, or 2a n + 2-e n - 2 n . 0, or 2-e . 2 - 2a, implying -e . log2(2 - 2a), or e . -log2(2 - 2a). The inequality above implies a . 1 (otherwise, 2 - 2a is negative) but a is also positive (since we assumed compression of symbol 1). The possible range of values of a is therefore (0, 1], and in this range e is greater than a, proving the statement above. (It is easy to see that a = 1 › e . -log20 = ., and a = 0.1 › e . -log2(2 - 20.1) . 0.10745.) It can be shown that this is just a special case of a general result that says, given an alphabet of n symbols, if we compress some of them by a certain factor, then the others must be expanded by a greater factor. One of my most productive days was throwing away 1000 lines of code. —Kenneth Thompson 72 2. Basic VL Codes 2.6 Tunstall Code The main advantage of variable-length codes is their variable lengths. Some codes are short, others are long, and a clever assignment of codes to symbols can produce compression. On the downside, variable-length codes are difficult to work with. The encoder has to construct each code from individual bits and pieces, has to accumulate and append several such codes in a short buffer, wait until n bytes of the buffer are full of code bits (where n must be at least 1), write the n bytes onto the output, shift the buffer n bytes, and keep track of the location of the last bit placed in the buffer. The decoder has to go through the reverse process. It is definitely easier to deal with fixed-length codes, and the Tunstall codes described here are an example of how such codes can be designed. The idea is to construct a set of fixed-length codes, each encoding a variable-length string of input symbols. As a result, these codes are also known as variable-to-fixed (or variable-to-block) codes, in contrast to the variable-length codes which are also referred to as fixed-to-variable. Imagine an alphabet that consists of two symbols A and B where A is more common. Given a typical string from this alphabet, we expect substrings of the form AA, AAA, AB, AAB, and B, but rarely strings of the form BB. We can therefore assign fixedlength codes to the following five substrings as follows. AA = 000, AAA = 001, AB = 010, ABA = 011, and B = 100. A rare occurrence of two consecutive Bs will be encoded by 100100, but most occurrences of B will be preceded by an A and will be coded by 010, 011, or 100. This example is both bad and inefficient. It is bad, because AAABAAB can be encoded either as the four codes AAA, B, AA, B or as the three codes AA, ABA, AB; encoding is not unique and may require several passes to determine the shortest code. This happens because our five substrings don’t satisfy the prefix property. This example is inefficient because only five of the eight possible 3-bit codewords are used. An n-bit Tunstall code should use all 2n codewords. Another point is that our codes were selected without considering the relative frequencies of the two symbols, and as a result we cannot be certain that this is the best code for our alphabet. Thus, an algorithm is needed to construct the best n-bit Tunstall code for a given alphabet of N symbols, and such an algorithm is given in [Tunstall 67]. Given an alphabet of N symbols, we start with a code table that consists of the symbols. We then iterate as long as the size of the code table is less than or equal to 2n (the number of n-bit codes). Each iteration performs the following steps: Select the symbol with largest probability in the table. Call it S. Remove S and include the N substrings Sx where x goes over all the N symbols. This step increases the table size by N - 1 symbols (some of them may be substrings). Thus, after iteration k, the table size will be N + k(N - 1) elements. If N + (k + 1)(N - 1) . 2n, perform another iteration (iteration k + 1). It is easy to see that the elements (symbols and substrings) of the table satisfy the prefix property and thus ensure unique encodability. If the first iteration adds element AB to the table, it must have removed element A. Thus, A, the prefix of AB, is not a code. If the next iteration creates element ABR, then it has removed element AB, 2.6 Tunstall Code 73 so AB is not a prefix of ABR. This construction also minimizes the average number of bits per alphabet symbol because of the requirement that each iteration select the element (or an element) of maximum probability. This requirement is similar to the way a Huffman code is constructed, and we illustrate it by an example. (a) 0.343 0.098 0.049 0.14 0.07 0.2 0.1 A A A B C B C B C (b) (c) 0.49 0.14 0.07 0.2 0.1 A A B C B C 0.7 0.2 0.1 A B C Figure 2.4: Tunstall Code Example. Given an alphabet with the three symbols A, B, and C (N = 3), with probabilities 0.7, 0.2, and 0.1, respectively, we decide to try to construct a set of 3-bit Tunstall codes (thus, n = 3). We start our code table as a tree with a root and three children (Figure 2.4a). In the first iteration, we select A and turn it into the root of a subtree with children AA, AB, and AC with probabilities 0.49, 0.14, and 0.07, respectively (Figure 2.4b). The largest probability in the tree is that of node AA, so the second iteration converts it to the root of a subtree with nodes AAA, AAB, and AAC with probabilities 0.343, 0.098, and 0.049, respectively (Figure 2.4c). After each iteration we count the number of leaves of the tree and compare it to 23 = 8. After the second iteration there are seven leaves in the tree, so the loop stops. Seven 3-bit codes are arbitrarily assigned to elements AAA, AAB, AAC, AB, AC, B, and C. The eighth available code should be assigned to a substring that has the highest probability and also satisfies the prefix property. The average bit length of this code is easily computed as 3 3(0.343 + 0.098 + 0.049) + 2(0.14 + 0.07) + 0.2 + 0.1 = 1.37 bits/symbol. In general, let pi and li be the probability and length of tree node i. If there are m nodes in the tree, the average bit length of the Tunstall code is n/m i=1 pili. The entropy of our alphabet is -(0.7Χlog2 0.7+0.2Χlog2 0.2+0.1Χlog2 0.1) = 1.156, so the Tunstall codes do not provide the best compression. The tree of Figure 2.4 is referred to as a parse tree, not a code tree. It is complete in the sense that every interior node has N children. Notice that the total number of nodes of this tree is 3 Χ 2 + 1 and in general a(N - 1) + 1. A parse tree defines a set of substrings over the alphabet (seven substrings in our example) such that any string of symbols from the alphabet can be broken up (subdivided) into these substrings (except that the last part may be only a prefix of such a substring) in one way only. The subdivision is unique because the set of substrings defined by the parse tree is proper, i.e., no substring is a prefix of another substring. An important property of the Tunstall codes is their reliability. If one bit becomes corrupt, only one code will get bad. Normally, variable-length codes are not robust. One 74 2. Basic VL Codes bad bit may corrupt the decoding of the remainder of a long sequence of such codes. It is possible to incorporate error-control codes in a string of variable-length codes, but this increases its size and reduces compression. Section 5.2.1 illustrates how a combination of the Tunstall algorithm with Huffman coding can improve compression in a two-step, dual tree process. A major downside of the Tunstall code is that both encoder and decoder have to store the complete code (the set of substrings of the parse tree). There are 10 types of people in this world: those who understand binary and those who don’t. —Author unknown 2.7 Schalkwijk’s Coding One of the many contributions of Isaac Newton to mathematics is the well-known binomial theorem. It states (a + b)n = n i=0 n i aibn-i, where the term n i = n! i!(n - i)! is pronounced “n over i” and is referred to as a binomial coefficient. Blaise Pascal (Section 3.24), a contemporary of Newton, discovered an elegant way to compute these coefficients without the lengthy calculations of factorials, multiplications, and divisions. He came up with the famous triangle that now bears his name (Figure 2.5) and showed that the general element of this triangle is a binomial coefficient. 1 1 1 1 2 1 1 3 3 1 1 4 6 4 1 1 5 10 10 5 1 1 6 15 20 15 6 1 1 7 21 35 35 21 7 1 1 8 28 56 70 56 28 8 1 1 9 36 84 126 126 84 36 9 1 1 10 45 120 210 252 210 120 45 10 1 1 11 55 165 330 462 462 330 165 55 11 1 1 12 66 220 495 792 924 792 495 220 66 12 1 1 13 78 286 715 1287 1716 1716 1287 715 286 78 13 1 1 14 91 364 1001 2002 3003 3432 3003 2002 1001 364 91 14 1 1 15 105 455 1365 3003 5005 6435 6435 5005 3003 1365 455 105 15 1 1 16 120 560 1820 4368 8008 11440 12870 11440 8008 4368 1820 560 120 16 1 Figure 2.5: Pascal Triangle. Quand on voit le style naturel, on est tout `etonn`e et ravi, car on s’attendait de voir un auteur, et on trouve un homme. (When we see a natural style, we are quite surprised and delighted, for we expected to see an author and we find a man.) —Blaise Pascal, Pens`ees (1670) 2.7 Schalkwijk’s Coding 75 The Pascal triangle is an infinite triangular matrix that’s constructed from the edges inwards. First fill the left and right edges with 1’s, then compute each interior element as the sum of the two elements directly above it. The construction is simple and it is trivial to derive an explicit expression for the general element of the triangle and show that it is a binomial coefficient. Number the rows from 0 starting at the top, and number the columns from 0 starting on the left. A general element is denoted by i j . Now observe that the top two rows (corresponding to i = 0 and i = 1) consist of 1’s and that every other row can be obtained as the sum of its predecessor and a shifted version of its predecessor. For example, 1 3 3 1 + 1 3 3 1 1 4 6 4 1 This shows that the elements of the triangle satisfy i 0 = i i = 1, i= 0, 1, . . . , i j = i - 1 j - 1 + i - 1 j , i= 2, 3, . . . , j = 1, . . . , (i - 1). From this, it is easy to obtain the explicit expression i j = i - 1 j - 1 + i - 1 j = (i - 1)! (j - 1)!(i - j)! + (i - 1)! j!(i - 1 - j)! = j(i - 1)! j!(i - j)! + (i - j)(i - 1)! j!(i - j)! = i! j!(i - j)! . And this is Newton’s binomial coefficient i j . The Pascal triangle has many interesting and unexpected properties, some of which are listed here. The sum of the elements of row i is 2i. If the second element of a row is a prime number, all the elements of the row (except the 1’s) are divisible by it. For example, the elements 7, 21, and 35 of row 7 are divisible by 7. Select any diagonal and any number of consecutive elements on it. Their sum will equal the number on the row below the end of the selection and off the selected diagonal. For example, 1 + 6 + 21 + 56 = 84. 76 2. Basic VL Codes Select row 7, convert its elements 1, 7, 21, 35, 35, 21, 7, and 1 to the single number 19487171 by concatenating the elements, except that a multidigit element is first carried over, such that 1, 7, 21, 35,. . . become 1(7+2)(1+3)(5+3) . . . = 1948 . . .. This number equals 117 and this magic-11 property holds for any row. The third column from the right consists of the triangular numbers 1, 3, 6, 10,. . . . Select all the odd numbers on the triangle and fill them with black. The result is the Sierpinski triangle (a well-known fractal). Other unusual properties can be found in the vast literature that exists on the Pascal triangle. The following is a quotation from Donald Knuth: “There are so many relations in Pascal’s triangle, that when someone finds a new identity, there aren’t many people who get excited about it anymore, except the discoverer.” The Pascal triangle is the basis of the unusual coding scheme described in [Schalkwijk 72]. This method starts by considering all the finite bit strings of length n that have exactly w 1’s. The set of these strings is denoted by T(n,w). If a string t consists of bits t1 through tn, then we define weights w1 through wn as the partial sums wk = n i=k ti. Thus, if t = 010100, then w1 = w2 = 2, w3 = w4 = 1, and w5 = w6 = 0. Notice that w1 always equals w. We now define, somewhat arbitrarily, a ranking i(t) on the n w strings in set T(n,w) by i(t) = n k=1 tkn - k wk . (The binomial coefficient i j is defined only for i . j, so we set it to 0 when i < n.) The rank of t = 010100 becomes 0 + 16 - 2 2 + 0 + 16 - 4 1 + 0 + 0 = 6 + 2 = 8. It can be shown that the rank of a string t in set T(n,w) is between 0 and n w - 1. The following table lists the ranking for the 6 2 = 15 strings of set T(6, 2). 0 000011 3 001001 6 010001 9 011000 12 100100 1 000101 4 001010 7 010010 10 100001 13 101000 2 000110 5 001100 8 010100 11 100010 14 110000 The first version of the Schalkwijk coding algorithm is not general. It is restricted to data symbols that are elements t of T(n,w). We assume that both encoder and decoder know the values of n and w. The method employs the Pascal triangle to determine the rank i(t) of each string t, and this rank becomes the code of t. The maximum value 2.7 Schalkwijk’s Coding 77 of the rank is n w - 1, so it can be expressed in log2 n w  bits. Thus, this method compresses each n-bit string (of which w bits are 1’s) to log2 n w  bits. Consider a source of bits that emits a 0 with probability q and a 1 with probability p = 1- q. The entropy of this source is H(p) = -p log2 p - (1 - p) log2(1 - p). In our strings, p = w/n, so the compression performance of this method is measured by the ratio log2 n w  n and it can be shown that this ratio approaches H(w/n) when n becomes very large. (The proof employs the famous Stirling formula n! . .2.n nne-n.) Figure 2.6a illustrates the operation of both encoder and decoder. Both know the values of n and w and they construct in the Pascal triangle a coordinate system tilted as shown in the figure and with its origin at element w of row n of the triangle. 1 1 1 1 2 1 1 3 3 1 1 4 6 4 1 1 5 10 10 5 1 1 6 15 20 15 6 1 x x y y 1 1 2 2 3 4 5 6 0 0 1 1 1 1 2 1 1 3 3 1 1 4 6 4 1 1 5 10 10 5 1 1 6 15 20 15 6 1 (a) (b) Figure 2.6: First Version. As an example, suppose that (n,w) = (6, 2). This puts the origin at the element 15 as shown in part (a) of the figure. The encoder starts at the origin, reads bits from the input string, and moves one step in the x direction for each 0 read and one step in the y direction for each 1 read. In addition, before moving in the y direction, the encoder saves the next triangle element in the x direction (the one it will not go to). Thus, given the string 010100, the encoder starts at the origin (15), moves to 10, 4, 3, 1, 1, and 1, while saving the values 6 (before it moves from 10 to 4) and 2 (before it moves from 3 to 1). The sum 6 + 2 = 8 = 10002 is the 4-bit rank of the input string and it becomes the encoder’s output. The decoder also knows the values of n and w, so it constructs the same coordinate system in the triangle and starts at the origin. Given the 4-bit input 10002 = 8, the decoder compares it to the next x value 10, and finds that 8 < 10. It therefore moves in the x direction, to 10, and emits a 0. The input 8 is compared to the next x value 6, but it is not less than 6. The decoder responds by subtracting 8 - 6 = 2, moving in the y direction, to 4, and emitting a 1. The current input, 2, is compared to the next x value 3, and is found to be smaller. The decoder therefore moves in the x direction, to 3, and emits a 0. When the input 2 is compared to the next x value 2, it is not smaller, so the decoder: (1) subtracts 2-2 = 0, (2) moves in the y direction to 1, and (3) emits a 1. The decoder’s input is now 0, so the decoder finds it smaller than the values on the x axis. It therefore keeps moving in the x direction, emitting two more zeros until it reaches the top of the triangle. 78 2. Basic VL Codes A similar variant is shown in Figure 2.6b. The encoder always starts at the apex of the triangle, moves in the -x direction for each 0 and in the -y direction for each 1, where it also records the value of the next element in the -x direction. Thus, the two steps in the -y direction in the figure record the values 1 and 3, whose sum 4 becomes the encoded value of string 010100. The decoder starts at 15 and proceeds in the opposite direction toward the apex. It is not hard to see that it ends up decoding the string 001010, which is why the decoder’s output in this variant has to be reversed before it is used. This version of Schalkwijk coding is restricted to certain bit strings, and is also block-to-block coding. Each block of n bits is replaced by a block of log2 n w  bits. The next version is similar, but is variable-to-block coding. We again assume a source of bits that emits a 0 with probability q and a 1 with probability p = 1- q. A string of n bits from this source may often have close to pn 1’s and qn zeros, but may sometimes have different numbers of zeros and 1’s. We select a convenient value for n, a value that is as large as possible and where both pn and qn are integers or very close to integers. If p = 1/3, for example, then n = 12 may make sense, because it results in np = 4 and nq = 8. We again employ the Pascal triangle and take a rectangular block of (pn + 1) rows and (qn + 1) columns such that the top of the triangle will be at the top-right corner of the rectangle (Figure 2.7). 1 1 1 1 1 1 1 1 8 7 6 5 4 3 2 1 36 28 21 15 10 6 3 1 120 84 56 35 20 10 4 1 y x pn 0 qn 1 2 3 1 B A C 2 3 4 5 6 7 Figure 2.7: Second Version. As before, we start at the bottom-left corner of the array and read bits from the source. For each 0 we move a step in the x direction and for each 1 we move in the y direction. If the next n bits have exactly pn 1’s, we will end up at point “A,” the top-right corner of the array, and encode n bits as before. If the first n bits happen to have more than pn 1’s, then the top of the array will be reached (after we have read np 1’s) early, say at point “B,” before we have read n bits. We cannot read any more source bits, because any 1 would take us outside the array, so we append several dummy zeros to what we have read, to end up with n bits (of which np are 1’s). This is encoded as before. Notice that the decoder can mimic this operation. It operates as before, but stops decoding when it reaches the top boundary. If the first n bits happen to have many zeros, the encoder will end up at the right boundary of the array, say, at point “C,” after it has read qn zeros but before it has read n bits. In such a case, the encoder appends several 1’s, to end up with exactly n bits (of which precisely pn are 1’s), and encodes as before. The decoder can mimic this operation by simply stopping when it reaches the right boundary. Any string that has too many or too few 1’s degrades the compression, because it encodes fewer than n bits in the same log2 n pn  bits. Thus, the method may not be very effective, but it is an example of a variable-to-block encoding. 2.8 Tjalkens–Willems V-to-B Coding 79 The developer of this method points out that the method can be modified to employ the Pascal triangle for block-to-variable coding. The value of n is determined and it remains fixed. Blocks of n bits are encoded and each block is preceded by the number of 1’s it contains. If the block contains w 1’s, it is encoded by the appropriate part of the Pascal triangle. Thus, each block of n bits may be encoded by a different part of the triangle, thereby producing a different-length code. The decoder can still work in lockstep with the encoder, because it first reads the number w of 1’s in a block. Knowing n and w tells it what part of the triangle to use and how many bits of encoding to read. It has been pointed out that this variant is similar to the method proposed by [Lynch 66] and [Davisson 66]. This variant has also been extended by [Lawrence 77], whose block-to-variable coding scheme is based on a Pascal triangle where the boundary points are defined in a special way, based on the choice of a parameter S. 2.8 Tjalkens–Willems V-to-B Coding The little-known variable-to-block coding scheme presented in this section is due to [Tjalkens and Willems 92] and is an extension of earlier work described in [Lawrence 77]. Like the Schalkwijk’s codes of Section 2.7 and the Lawrence algorithm, this scheme employs the useful properties of the Pascal triangle. The method is based on the choice of a positive integer parameter C. Once a value for C has been selected, the authors show how to construct a set L of M variable-length bitstrings that satisfy the following: 1. Set L is complete. Given any infinite bitstring (in practice, a string of M or more bits), L contains a prefix of the string. 2. Set L is proper. No segment in the set is a prefix of another segment. Once L has been constructed, it is kept in lexicographically-sorted order, so each string in L has an index between 0 and M - 1. The input to be encoded is a long bitstring. It is broken up by the encoder into segments of various lengths that are members of L. Each segment is encoded by replacing it with its index in L. Note that the index is a (log2M)-bit number. Thus, if M = 256, each segment is encoded in eight bits. The main task is to construct set L in such a way that the encoder will be able to read the input bit by bit, stop when it has read a bit pattern that is a string in L, and determine the code of the string (its index in L). The theory behind this method is complex, so only the individual steps and tests are summarized here. Given a string s of a zeros and b 1’s, we define the function Q(s) = (a + b + 1)a + b b . (The authors show that 1/Q is the probability of string s.) We denote by s-1 the string s without its last (rightmost) bit. String s is included in set L if it satisfies Q(s-1) < C . Q(s). (2.4) The authors selected C = 82 (because this results in the convenient size M = 256). Once C is known, it is easy to decide whether a given string s with a zeros and b 1’s is a 80 2. Basic VL Codes member of set L (i.e., whether s satisfies Equation (2.4)). If s is in L, then point (a, b) in the Pascal triangle (i.e., element b of row a, where row and column numbering starts at 0) is considered a boundary point. Figure 2.8a shows the boundary points (underlined) for C = 82. 1 12 12 1 1 11 11 1 1 10 10 1 1 9 36 36 9 1 1 8 28 28 8 1 1 7 21 21 7 1 1 6 15 20 15 6 1 1 5 10 10 5 1 4 6 4 1 1 1 3 3 1 1 2 1 1 1 1 1 81 81 1 a b 70 1 1 70 71 1 1 71 72 1 1 72 73 1 1 1 1 73 74 2 1 1 2 74 76 3 1 1 3 76 79 4 1 1 1 4 79 83 5 2 2 5 88 7 4 7 88 83 95 11 11 95 106 22 106 128 128 256 1 1 1 1 a b (a) (b) Figure 2.8: (a) Boundary Points. (b) Coding Array. The inner parts of the triangle are not used in this method and can be removed. Also, The lowest boundary points are located on row 81 and lower parts of the triangle are not used. If string s is in set L, then we can start at the apex of the triangle, move in the a direction for each 0 and in the b direction for each 1 in s, and end up at a boundary point. The figure illustrates this walk for the string 0010001, where the boundary point reached is 21. Setting up the partial Pascal triangle of Figure 2.8a is just the first step. The second step is to convert this triangle to a coding triangle M(a, b) of the same size, where each walk for a string s can be used to determine the index of s in set L and thus its code. The authors show that element M(a, b) of this triangle must equal the number of distinct ways to reach a boundary point after processing a zeros and b 1’s (i.e., after moving a steps in the a direction and b steps in the b direction). Triangle M(a, b) is constructed according to M(a, b) = 1, if (a, b) is a boundary point, M(a + 1, b) +M(a, b + 1), otherwise. The coding array for C = 82 (constructed from the bottom up) is shown in Figure 2.8b. Notice that its apex, M(0, 0), equals the total number of strings in L. Once this triangle is available, both encoding and decoding are simple and are listed in Figure 2.9a,b. The former inputs individual bits and moves inM(a, b) in the a or b directions according to the inputs. The end of the current input string is signalled when a node with a 1 is reached in the coding triangle. For each move in the b direction, the next element in the a direction (the one that will not be reached) is added to the index. At 2.9 Phased-In Codes 81 the end, the index is the code of the current string. Figure 2.8b shows the moves for 0010001 and how the nodes 95 and 3 are selected and added to become code 98. The decoder starts with the code of a string in variable index. It compares index to the sum (I +M(a + 1, b)) and moves in the a or b directions according to the result, generating one output bit as it moves. Decoding is complete when the decoder reaches a node with a 1. index:=0; a:=0; b:=0; while M(a, b) = 1 do if next input = 0 then a:=a+1 else index:=index+M(a + 1, b); b:=b+1 endif endwhile I:=0; a:=0; b:=0; while M(a, b) = 1 do if index < (I +M(a + 1, b)) then next_output:=0; a:=a+1; else next_output:=1; I := I+M(a+1, b); b:=b+1 endif endwhile Figure 2.9: (a) Encoding and (b) Decoding. Extraordinary how mathematics help you. . . . —Samuel Beckett, Molloy (1951) 2.9 Phased-In Codes Many of the prefix codes described here were developed for the compression of specific types of data. These codes normally contain codewords of various lengths and they are suitable for the compression of data where individual symbols have widely different probabilities. Data where symbols have equal probabilities cannot be compressed by VLCs and is normally assigned fixed-length codes. The codes of this section (also called phased-in binary codes, see Appendix A-2 in [Bell et al. 90]) constitute a compromise. A set of phased-in codes consists of codewords of two lengths and may contribute something (although not much) to the compression of data where symbols have roughly equal probabilities. Section 5.3.1 and page 376 discuss applications of these codes. Given n data symbols, where n = 2m (implying that m = log2 n), we can assign them m-bit codewords. However, if 2m-1 < n < 2m, then log2 n is not an integer. If we assign fixed-length codes to the symbols, each codeword would be log2 n bits long, but not all the codewords would be used. The case n = 1,000 is a good example. In this case, each fixed-length codeword is log2 1,000 = 10 bits long, but only 1,000 out of the 1,024 possibe codewords are used. In the approach described here, we try to assign two sets of codes to the n symbols, where the codewords of one set are m - 1 bits long and may have several prefixes and the codewords of the other set are m bits long and have different prefixes. The average length of such a code is between m - 1 and m bits and is shorter when there are more short codewords. 82 2. Basic VL Codes A little research and experimentation leads to the following method of constructing the two sets of codes. Given a set of n data symbols (or simply the integers 0 through n - 1) where 2m . n < 2m+1 for some integer m, we denote n = 2m + p where 0 . p < 2m - 1 and also define P def = 2m - p. We construct 2p long (i.e., m-bit) codewords and P short, (m - 1)-bit codewords. The total number of codewords is always 2p + P = 2p + 2m - p = (n - 2m) + 2m = n and there is an even number of long codewords. The first P integers 0 through P - 1 receive the short codewords and the remaining 2p integers P through n - 1 are assigned the long codewords. The short codewords are the integers 0 through P - 1, each encoded in m - 1 bits. The long codewords consist of p pairs, where each pair starts with the m-bit value P, P +1, . . . , P +p-1, followed by an extra bit, 0 or 1, to distinguish between the two codewords of a pair. The following tables illustrate both the method and its compression performance. Table 2.10 lists the values of m, p, and P for n = 7, 8, 9, 15, 16, and 17. Table 2.11 lists the actual codewords. n m p= n - 2m P = 2m - p 7 2 3 1 8 3 0 8 9 3 1 7 15 3 7 1 16 4 0 16 17 4 1 15 Table 2.10: Parameters of Phased-In Codes. i n=7 8 9 15 16 17 0 00 000 000 000 0000 0000 1 010 001 001 0010 0001 0001 2 011 010 010 0011 0010 0010 3 100 011 011 0100 0011 0011 4 101 100 100 0101 0110 0100 5 110 101 101 0110 0101 0101 6 111 110 110 0111 0110 0110 7 111 1110 1000 0111 0111 8 1111 1001 1000 1000 9 1010 1001 1001 10 1011 1010 1010 11 1100 1011 1011 12 1101 1100 1100 13 1110 1101 1101 14 1111 1110 1110 15 1111 11110 16 11111 Table 2.11: Six Phased-In Codes. 2.9 Phased-In Codes 83 Table 2.11 is also the key to estimating the compression ratio produced by these codes. The table shows that when n approaches a power of 2 (such as n = 7 and n = 15), there are few short codewords and many long codewords, indicating low efficiency. When n is a power of 2, all the codewords are long and the method does not produce any compression (the compression ratio is 1). When n is slightly greater than a power of 2 (such as n = 9 and 17), there are many short codewords and therefore better compression. The total size of the n phased-in codewords for a given n is P Χm+ 2p(m+ 1) = (2m+1 - n)m+ 2(n - 2m)(m+ 1) where m = log2 n, whereas the ideal total size of n fixed-length codewords (keeping in mind the fact that log2 n is normally a noninteger), is nlog2 n. Thus, the ratio of these two quantities is the compression ratio of the phased-in codes. It is illustrated in Figure 2.12 for n values 2 through 256. 50 100 150 200 250 0.85 0.90 0.95 1.00 m:=Floor[Log[2,n]]; Plot[(n m+2 n-2^(m+1))/(n Ceiling[ Log[2,n]]), {n,2,256}] Figure 2.12: Efficiency of Phased-In Codes. It is obvious that the compression ratio goes up toward 1 (indicating worse compression) as n reaches a power of 2, then goes down sharply (indicating better compression) as we pass that value. See http://f-cpu.seul.org/whygee/phasing-in_codes/phasein.html for other examples and a different way to measure the efficiency of these codes. Another advantage of phased-in codes is the ease of encoding and decoding them. Once the encoder is given the value of n, it computes m, p, and P. Given an integer i to be encoded, if it is less than P, the encoder constructs an m-bit codeword with the value i. Otherwise, the encoder constructs an (m + 1)-bit codeword where the first m bits (the prefix) have the value P + (i - P) χ 2 and the rightmost bit is (i - P) mod 2. Using n = 9 as an example, we compute m = 3, p = 1, and P = 7. If i = 6, then 84 2. Basic VL Codes i < P, so the 3-bit codeword 110 is prepared. If i = 8, then the prefix is the three bits 7 + (8 - 7) χ2 = 7 = 1112 and the rightmost bit is (8 - 7) mod 2 = 1. The decoder also starts with the given value of n and computes m, p, and P. It inputs the next m bits into i. If this value is less than P, then this is the entire codeword and it is returned by the decoder as the result. Otherwise, the decoder inputs the next bit b and appends it to 2(i - P) + P. Using n = 9 as an example, if the decoder inputs the three bits 111 (equals P) into i, it inputs the next bit into b and appends it to 2(7 - 7) + 7 = 1112, resulting in either 1110 or 1111, depending on b. The phased-in codes are closely related to the minimal binary code of Section 3.13. See [seul.org 06] for a Mathematica notebook to construct phased-in codes. Two variations: In the basic phased-in codes algorithm, the shorter codes are assigned to the first symbols. This makes sense if it is known (or suspected) that these symbols are more common. In some cases, however, we may know that the middle symbols are the most common, and such cases can benefit from rotating the original sequence of phased-in codes such that the shorter codes are located in the middle. This variation is termed centered phased-in codes. Table 2.11 shows that for n = 9 there are seven 3-bit codes and two 4-bit codes. The sequence of code lengths is therefore (3, 3, 3, 3, 3, 3, 3, 4, 4) and it can be rotated one position to the right to produce (4, 3, 3, 3, 3, 3, 3, 3, 4). The FELICS image compression algorithm (Section 7.23) is an example of common symbols in the middle of a range. The values in the middle region of part (b) of Figure 7.133 are more common than their extreme neighbors and should therefore be assigned shorter codes. Another useful version, the reverse centered phased-in codes, is obtained when the longer codes are placed in the middle. Rotating the sequence of code lengths (3, 3, 3, 3, 3, 3, 3, 4, 4) three places to the left results in (3, 3, 3, 3, 4, 4, 3, 3, 3). It is also possible to construct suffix phased-in codes. We start with a set of fixedlength codes and convert it to two sets of codewords by removing the leftmost bit of some codewords. This bit is removed if it is a 0 and if its removal does not create any ambiguity. Table 2.13 (where the removed bits are in italics) illustrates an example for the first 24 nonnegative integers. The fixed-length representation of these integers requires five bits, but each of the eight integers 8 through 15 can be represented by only four bits because 5-bit codes can represent 32 symbols and we have only 24 symbols. A simple check verifies that, for example, coding the integer 8 as 1000 instead of 01000 does not introduce any ambiguity, because none of the other 23 codes ends with 1000. One-third of the codewords in this example are one bit shorter, but if we consider only the 17 integers from 0 to 16, about half will require four bits instead of five. The efficiency of this code depends on where n (the number of symbols) is located in the interval [2m, 2m+1 - 1). 00000 00001 00010 00011 00100 00101 00110 00111 01000 01001 01010 01011 01100 01101 01110 01111 10000 10001 10010 10011 10100 10101 10110 10111 Table 2.13: Suffix Phased-In Codes. 2.10 Redundancy Feedback (RF) Coding 85 The suffix phased-in codes are suffix codes (if c has been selected as a codeword, no other codeword will end with c). Suffix codes can be considered the complements of prefix codes and are also mentioned in Section 4.5. 2.10 Redundancy Feedback (RF) Coding The interesting and original method of redundancy feedback (RF) coding is the brainchild of Eduardo Enrique Gonz΄alez Rodr΄ύguez who hasn’t published it formally. As a result, information about it is hard to find. At the time of writing (early 2007), there is a discussion in file new-entropy-coding-algorithm-312899.html at web site http://archives.devshed.com/forums/compression-130/and some information (and source code) can also be obtained from the authors of this book. The method employs phased-in codes, but is different from other entropy coders. It may perhaps be compared with static arithmetic coding. Most entropy coders assign variable-length codes to data symbols such that the length of the code for a symbol is inversely proportional to the symbol’s frequency of occurrence. The RF method, in contrast, starts by assigning several fixed-length (block) codes to each symbol according to its probability. The method then associates a phased-in code (that the developer terms “redundant information”) with each block codeword. Encoding is done in reverse, from the end of the input stream. Each symbol is replaced by one of its block codes B in such a way that the phased-in code associated with B is identical to some bits at the start (the leftmost part) of the compressed stream. Those bits are deleted from the compressed stream (which generates compression) and B is prepended to it. For example, if the current block code is 010111 and the compressed stream is 0111|0001010 . . ., then the result of prepending the code and removing identical bits is 01|0001010 . . .. We start with an illustrative example. Given the four symbols A, B, C, and D, with probabilities 37.5%, 25%, 12.5%, and 25%, respectively, we assign each symbol several block codes according to its probability. The total number of codes must be a power of 2, so A is assigned three codes, each of B and D gets two codes, and C becomes the “owner” of one code, for a total of eight codes. Naturally, the codes are the 3-bit numbers 0 through 7. Table 2.14a lists the eight codes and their redundant information (the associated phased-in codes). Thus, e.g., the three codes of A are associated with the phased-in codes 0, 10, and 11, because these are the codes for n = 3. (Section 2.9 shows that we have to look for the integer m that satisfies 2m . n < 2m+1. Thus, for n = 3, m is 1. The first 2m = 2 symbols are assigned the 2-bit numbers 0 + 2 and 1 + 2 and the remaining 3-2 symbol is assigned the 1-bit number i-2m = 2-2 = 0.) Similarly, the two phased-in codes associated with B are 0 and 1. Symbol D is associated with the same two codes, and the single block code 5 of C has no associated phased-in codes because there are no phased-in codes for a set of one symbol. Table 2.14b is constructed similarly and lists 16 4-bit block codes and their associated phased-in codes for the three symbols A, B, and C with probabilities 0.5, 0.2, and 0.2, respectively. First, a few words on how to determine the number of codes per symbol from the number n of symbols and their frequencies fi. Given an input string of F symbols (from an alphabet of n symbols) such that symbol i appears fi times (so that fi = F), we first determine the number of codes. This is simply the power m of 2 that 86 2. Basic VL Codes satisfies 2m-1 < n . 2m. We now multiply each fi by 2m/F. The new sum satisfies fi Χ 2m/F = 2m. Next, we round each term of this sum to the nearest integer, and if any is rounded down to zero, we set it to 1. Finally, if the sum of these integers is slightly different from 2m, we increment (or decrement) each of the largest ones by 1 until the sum equals 2m. Code Symbol Redundant Info 0 A 0/3 › 0 1 A 1/3 › 10 2 A 2/3 › 11 3 B 0/2 › 0 4 B 1/2 › 1 5 C 0/1›- 6 D 0/2 › 0 7 D 1/2 › 1 Code Symbol Redundant Info 0 0000 A 0/10 › 000 1 0001 A 1/10 › 001 2 0010 A 2/10 › 010 3 0011 A 3/10 › 011 4 0100 A 4/10 › 100 5 0101 A 5/10 › 101 6 0110 A 6/10 › 1100 7 0111 A 7/10 › 1101 8 1000 A 8/10 › 1110 9 1001 A 8/10 › 1111 10 1010 B 0/3 › 0 11 1011 B 1/3 › 10 12 1100 B 2/3 › 11 13 1101 C 0/3 › 0 14 1110 C 1/3 › 10 15 1111 C 2/3 › 11 (a) (b) Table 2.14: Eight and 16 RF Codes. As an example, consider a 6-symbol alphabet and an input string of F = 47 symbols, where the six symbols appear 17, 6, 3, 12, 1, and 8 times. We first determine that 22 < 6 . 23, so we need 8 codes. Multiplying 17Χ8/47 = 2.89 › 3, 6Χ8/47 = 1.02 › 1, 3Χ8/47 = 0.51 › 1, 12Χ8/47 = 2.04 › 2, 1Χ8/47 = 0.17 › 0, and 8Χ8/47 = 1.36 › 1. The last step is to increase the 0 to 1, and make sure the sum is 8 by decrementing the largest count, 3, to 2. The codes of Table 2.14b are now used to illustrate RF encoding. Assume that the input is the string AABCA. It is encoded from end to start. The last symbol A is easy to encode as we can use any of its block codes. We therefore select 0000. The next symbol, C, has three block codes, and we select 13 = 1101. The associated phased-in code is 0, so we start with 0000, delete the leftmost 0, and prepend 1101, to end up with 1101|000. The next symbol is B and we select block code 12 = 1100 with associated phased-in code 11. Encoding is done by deleting the leftmost 11 and prepending 1100, to end up with 1100|01|000. To encode the next A, we select block code 6 = 0110 with associated phased-in code 1100. Again, we delete 1100 and prepend 0110 to end up with 0110||01|000. Finally, the last (i.e., leftmost) symbol A is reached, for which we select block code 3 = 0011 (with associated phased-in code 011) and encode by deleting 011 and prepending 0011. The compressed stream is 0011|0||01|000. 2.10 Redundancy Feedback (RF) Coding 87 The RF encoding principle is simple. Out of all the block codes assigned to the current symbol, we select the one whose associated phased-in code is identical to the prefix of the compressed stream. This results in the deletion of the greatest number of bits and thus in maximum compression. Decoding is the opposite of encoding. The decoder has access to the table of block codes and their associated codes (or it starts from the symbols’ probabilities and constructs the table as the encoder does). The compressed stream is 0011001000 and the first code is the leftmost four bits 0011 = 3 › A. The first decoded symbol is A and the decoder deletes the 0011 and prepends 011 (the phased-in code associated with 3) to end up with 011001000. The rest of the decoding is straightforward. Experiments with this method verify that its performance is generally almost as good as Huffman coding. The main advantages of RF coding are as follows: 1. It works well with a 2-symbol alphabet. We know that Huffman coding fails in this situation, even when the probabilities of the symbols are skewed, because it simply assigns the two 1-bit codes to the symbols. In contrast, RF coding assigns the common symbol many (perhaps seven or 15) codes, while the rare symbol is assigned only one or two codes, thereby producing compression even in such a case. 2. The version of RF presented earlier is static. The probabilities of the symbols have to be known in advance in order for this version to work properly. It is possible to extend this version to a simple dynamic RF coding, where a buffer holds the most-recent symbols and code assignment is constantly modified. This version is described below. 3. It is possible to replace the phased-in codes with a simple form of arithmetic coding. This slows down both encoding and decoding, but results in better compression. Dynamic RF coding is slower than the static version above, but is more efficient. Assuming that the probabilities of the data symbols are unknown in advance, this version of the basic RF scheme is based on a long sliding buffer. The buffer should be long, perhaps 215 symbols or longer, in order to reflect the true frequencies of the symbols. A common symbol will tend to appear many times in the buffer and will therefore be assigned many codes. For example, given the alphabet A, B, C, D, and E, with probabilities 60%, 10%, 10%, 15%, and 5%, respectively, the buffer may, at a certain point in the encoding, hold the following raw ‹ A B A A B A C A A D A A B A A A C A A D A C A E A A A D A A C A A B A A 36 35 34 33 32 31 30 29 28 27 26 25 24 23 22 21 20 19 18 17 16 15 14 13 12 11 10 9 8 7 6 5 4 3 2 1 coded ‹ On the left, there is raw (unencoded) text and on the right there is text that has already been encoded. We can imagine the text being stationary and the buffer sliding to the left. If the buffer is long enough, the text inside it will reflect the true probabilities of the symbols and each symbol will have a number of codes proportional to its probability. At any time, the symbol immediately to the right of the buffer is encoded by selecting one of its codes in the buffer, and then moving the buffer one symbol to the left. If the buffer happens to contain no occurrences of the symbol to be encoded, then the code of all zeros is selected (which is why the codes in the buffer start at 1) and is output, followed by the raw (normally ASCII) code of the symbol. Notice that sliding the buffer modifies the codes of the symbols, but the decoder can do this in lockstep with the encoder. Once a code has been selected for a symbol, the code is prepended to the 88 2. Basic VL Codes compressed stream after its associated phased-in code is used to delete identical bits, as in the static version. We illustrate this version with a 4-symbol alphabet and the string ABBACBBBAADA. We assume a buffer with seven positions (so the codes are between 1 and 7) and place the buffer initially such that the rightmost A is immediately to its right, thus ABBA[CBBBAAD]A. The initial buffer position and codes (both the 3-bit RF codes and the associated phased-in codes) are shown here. Symbol A is immediately to the right of the buffer and it can be encoded as either 2 or 3. We arbitrarily select 2, ending up with a compressed stream of 010. A B B A C B B B A A D A 7 6 5 4 3 2 1 111 110 101 100 011 010 001 1:- 3:2 3:1 3:0 2:1 2:0 1:- - 11 10 0 1 0 - The buffer is slid, as shown below, thereby changing all the codes. This is why the dynamic version is slow. Symbol D is now outside the buffer and must be encoded as the pair (000, D) because there are no occurrences of D inside the buffer. The compressed stream becomes 000: 01000100|010. A B B A C B B B A A D A 7 6 5 4 3 2 1 111 110 101 100 011 010 001 3:2 1:- 3:2 3:1 3:0 3:1 3:0 11 - 11 10 0 10 0 Now comes another A that can be encoded as either 1 or 6. Selecting the 1 also selects its associated phased-in code 0, so the leftmost 0 is deleted from the compressed stream and 001 is prepended. The result is 001|00: 01000100|010. A B B A C B B B A A D A 7 6 5 4 3 2 1 111 110 101 100 011 010 001 4:3 2:1 1:- 4:2 4:1 4:0 2:0 11 1 - 10 01 00 0 The next symbol to be encoded is the third A from the right. The only available code is 5, which has no associated phased-in code. The output therefore becomes 101|001|00: 01000100|010. A B B A C B B B A A D A 7 6 5 4 3 2 1 111 110 101 100 011 010 001 5:4 5:3 1:- 1:- 5:2 5:1 5:0 111 110 - - 10 01 00 Next in line is the B. Four codes are available, of which the best choice is 5, with associated phased-in code 10. The string 101|1|001|00: 01000100|010 now becomes the current output. 2.11 Recursive Phased-In Codes 89 A B B A C B B B A A D A 7 6 5 4 3 2 1 111 110 101 100 011 010 001 2:1 4:3 4:2 2:0 1:- 4:1 4:0 1 11 10 0 - 01 00 Encoding continues in this way even though the buffer is now only partially full. The next B is encoded with only three Bs in the buffer, and with each symbol encoded, fewer symbols remain in the buffer. Each time a symbol s is encoded that has no copies left in the buffer, it is encoded as a pair of code 000 followed by the ASCII code of s. As the buffer gradually empties, more and more pairs are prepended to the output, thereby degrading the compression ratio. The last symbol (which is encoded with an empty buffer) is always encoded as a pair. Thus, the decoder starts with an empty buffer, and reads the first code (000) which is followed by the ASCII code of the first (leftmost) symbol. That symbol is shifted into the buffer, and decoding continues as the reverse of encoding. 2.11 Recursive Phased-In Codes The recursive phased-in codes were introduced in [Acharya and J΄aJ΄a 95] and [Acharya and J΄aJ΄a 96] as an enhancement to the well-known LZW (Lempel Ziv Welch) compression algorithm [Section 6.13]. These codes are easily explained in terms of complete binary trees, although their practical construction may be simpler with the help of certain recursive formulas conceived by Steven Pigeon. The discussion in Section 3.19 shows that any positive integer N can be written uniquely as the sum of certain powers of 2. Thus, for example, 45 is the sum 25 + 23 + 22 + 20. In general, we can write N = s i=1 2ai , where a1 > a2 > · · · > as . 0 and s . 1. For N = 45, for example, these values are s = 4 and a1 = 5, a2 = 3, a3 = 2, and a4 = 0. Once a value for N has been selected and the values of all its powers ai determined, a set of N variable-length recursive phased-in codes can be constructed from the tree shown in Figure 2.15. For each power ai, this tree has a subtree that is a complete binary tree of height ai. The individual subtrees are connected to the root by 2s - 2 edges labeled 0 or 1 as shown in the figure. a1 a2 a3 as T1 T2 T3 Ts complete complete complete binary binary tree tree 0 0 0 1 1 1 Figure 2.15: A Tree for Recursive Phased-In Codes. 90 2. Basic VL Codes The tree for N = 45 is shown in Figure 2.16. It is obvious that the complete binary tree for ai has ai leaves and the entire tree therefore has a total of N leaf nodes. The codes are assigned by sliding down the tree from the root to each leaf, appending a 0 to the code each time we slide to the left and a 1 each time we slide to the right. Some codes are also shown in the figure. Thus, the 45 recursive phased-in codes for N = 45 are divided into the four sets 0xxxxx, 10xxx, 110xx, and 111, where the x’s stand for bits. The first set consists of the 32 5-bit combinations prepended by a 0, the second set includes eight 5-bit codes that start with 10, the third set has four codes, and the last set consists of the single code 111. As we scan the leaves of each subtree from left to right, we find that the codewords in each set are in ascending order. Even the codewords in different sets appear sorted, when scanned from left to right, if we append enough zeros to each so they all become six bits long. The codes are prefix codes, because, for example, once a code of the form 0xxxxx has been assigned, no other codes will start with 0. 0 0 0 11001 10101 011110 1 1 1 111 Figure 2.16: A Tree for N = 45. In practice, these codes can be constructed in two steps, one trivial and the other one simple, as follows: 1. (This step is trivial.) Given a positive integer N, determine its powers ai. Given, for example, 45 = . . . 000101101 we first locate its leftmost 1. The position of this 1 (its distance from the right end) is s. We then scan the bits from right to left while decrementing an index i from s to 1. Each 1 found designates a power ai. 2. There is no need to actually construct any binary trees. We build the set of codes for a1 by starting with the prefix 0 and appending to it all the a1-bit numbers. The set of codes for a2 is similarly built by starting with the prefix 10 and appending to it all the a2-bit numbers. In his PhD thesis [Pigeon 01b], Steven Pigeon proposes a recursive formula as an alternative to the steps above. Following Elias (Section 3.4), we denote by ίk(n) the k-bit binary representation of the integer n. Given N, we find its largest power k, so N = 2k + b where 0 . b < 2k (k equals a1 above). The N recursive phased-in codes CN(n) for n = 0, 1, . . . , N - 1 are computed by CN(n) =... 0 : Cb(n), if 0 . n < b, ίk(n), if b = 0, 1 : ίk(n - b), otherwise. 2.11 Recursive Phased-In Codes 91 Their lengths LN(n) are given by LN(n) =... 1 + Lb(n), if 0 . n < b, k, if b = 0, 1 + k, otherwise. Table 2.17 lists the resulting codes for N = 11, 13, and 18. It is obvious that these are slightly different from the original codes of Acharya and J΄aJ΄a. The latter code for N = 11, for example, consists of the sets 0xxx, 10x, and 11, while Pigeon’s formula generates the sets 1xxx, 01x, and 00. n 11 13 18 0 00 00 00 1 010 0100 01 2 011 0101 10000 3 1000 0110 10001 4 1001 0111 10010 5 1010 1000 10011 6 1011 1001 10100 7 1100 1010 10101 8 1101 1011 10110 9 1110 1100 10111 10 1111 1101 11000 11 1110 11001 12 1111 11010 13 11011 14 11100 15 11101 16 11110 17 11111 Table 2.17: Codes for 11, 13, and 18. The recursive phased-in codes bear a certain resemblance to the start-step-stop codes of Section 3.2, but a quick glance at Table 3.3 shows the difference between the two types of codes. A start-step-stop code consists of sets of codewords that start with 0, 10, 110, and so on and get longer and longer, while the recursive phased-in codes consist of sets that start with the same prefixes but get shorter. “The grand aim of all science is to cover the greatest number of empirical facts by logical deduction from the smallest number of hypotheses or axioms,” he [Einstein] maintained. The same principle is at work in Ockham’s razor, in Feynman’s panegyric upon the atomic doctrine, and in the technique of data compression in information technology—all three of which extol economy of expression, albeit for different reasons. —Hans Christian von Baeyer, Information, The New Language of Science (2004) 92 2. Basic VL Codes 2.12 Self-Delimiting Codes Before we look at the main classes of VLCs, we list in this short section a few simple techniques (the first few of which are due to [Chaitin 66]) to construct self-delimiting codes, codes that have variable lengths and can be decoded unambiguously. 1. Double each bit of the original message, so the message becomes a set of pairs of identical bits, then append a pair of different bits. Thus, the message 01001 becomes the bitstring 00|11|00|00|11|01. This is simple but is obviously too long. It is also fragile, because one bad bit will confuse any decoder (computer or human). A variation of this technique precedes each bit of the number with an intercalary bit of 1, except the last bit, which is preceded with a 0. Thus, 01001 become ―10―11―10―10―01. We can also concentrate the intercalary bits together and have them followed by the number, as in 11110|01001 (which is the number itself preceded by its unary code). 2. Prepare a header with the length of the message and prepend it to the message. The size of the header depends on the size of the message, so the header should be made self-delimiting using method 1 above. Thus, the 6-bit message 010011 becomes the header 00|11|11|00|01 followed by 010011. It seems that the result is still very long (16 bits to encode six bits), but this is because our message is so short. Given a 1-million bit message, its length requires 20 bits. The self-delimiting header is therefore 42 bits long, increasing the length of the original message by 0.0042%. 3. If the message is extremely long (trillions of bits) its header may become too long. In such a case, we can make the header itself self-delimiting by writing it in raw format and preceding it with its own header, which is made self-delimiting with method 1. 4. It is now clear that there may be any number of headers. The first header is made self-delimiting with method 1, and all the other headers are concatenated to it in raw format. The last component is the (very long) original binary message. 5. A decimal variable-length integer can be represented in base 15 (quindecimal) as a string of nibbles (groups of four bits each), where each nibble is a base-15 digit (i.e., between 0 and 14) and the last nibble contains 16 = 11112. This method is sometimes referred to as nibble code or byte coding. Table 3.25 lists some examples. 6. A variation on the nibble code is to start with the binary representation of the integer n (or n - 1), prepend it with zeros until the total number of bits is divisible by 3, break it up into groups of three bits each, and prefix each group with a 0, except the leftmost (or alternatively, the rightmost) group, which is prepended by a 1. The length of this code for the integer n is 4(log2 n)/3, so it is ideal for a distribution of the form 2-4(log2 n)/3 . 1 n4/3 . (2.5) This is a power law distribution with a parameter of 3/4. A natural extension of this code is to k-bit groups. Such a code fits power law distributions of the form 1 n1+ 1 k . (2.6) 7. If the data to be compressed consists of a large number of small positive integers, then a word-aligned packing scheme may provide good (although not the best) compression combined with fast decoding. The idea is to pack several integers into fixed-length 2.12 Self-Delimiting Codes 93 fields of a computer word. Thus, if the word size is 32 bits, 28 bits may be partitioned into several k-bit fields while the remaining four bits become a selector that indicates the value of k. The method described here is due to [Anh and Moffat 05] who employ it to compress inverted indexes. The integers they compress are positive and small because they are differences of consecutive pointers that are in sorted order. The authors describe three packing schemes, of which only the first, dubbed simple-9, is discussed here. Simple-9 packs several small integers into 28 bits of a 32-bit word, leaving the remaining four bits as a selector. If the next 28 integers to be compressed all have values 1 or 2, then each can fit in one bit, making it possible to pack 28 integers in 28 bits. If the next 14 integers all have values of 1, 2, 3, or 4, then each fits in a 2-bit field and 14 integers can be packed in 28 bits. At the other extreme, if the next integer happens to be greater than 214 = 16, 384, then the entire 28 bits must be devoted to it, and the 32-bit word contains just this integer. The choice of 28 is particularly fortuitous, because 28 is divisible by 1, 2, 3, 4, 5, 7, 9, 14, and itself. Thus, a 32-bit word packed in simple-9 may be partitioned in nine ways. Table 2.18 lists these nine partitions and shows that at most three bits are wasted (in row e). Number Code Unused Selector of codes length bits a 28 1 0 b 14 2 0 c 9 3 1 d 7 4 0 e 5 5 3 f 4 7 0 g 3 9 1 h 2 14 0 i 1 28 0 Table 2.18: Summary of the Simple-9 Code. Given the 14 integers 4, 6, 1, 1, 3, 5, 1, 7, 1, 13, 20, 1, 12, and 20, we encode the first nine integers as c|011|101|000|000|010|100|000|110|000|b and the following five integers as e|01100|10011|00000|01011|10011|bbb, for a total of 64 bits, where each b indicates an unused bit. The originators of this method point out that the use of a Golomb code would have compressed the 14 integers into 58 bits, but the small loss of compression efficiency of simple-9 is often more than compensated for by the speed of decoding. Once the leftmost four bit of a 32-bit word are examined and the selector value is determined, the remaining 28 bits can be unpacked with a few simple operations. Allocating four bits for the selector is somewhat wasteful, because only nine of the 16 possible values are used, but the flexibility of the simple-9 code is the result of the many (nine) factors of 28. It is possible to give up one selector value, cut the selector size to three bits and increase the data segment to 29 bits, but 29 is a prime number, so a 29- bit segment cannot be partitioned into equal-length fields. The authors propose dividing a 32-bit word into a 2-bit selector and a 30-bit segment for packing data. The integer 94 2. Basic VL Codes 30 has 10 factors, so a table of the simple-10 code, similar to Table 2.18, would have 10 rows. The selector field, however, can specify only four different values, which is why the resulting code (not described here) is more complex and is denoted by relative-10 instead of simple-10. Intercalary: Inserted between other elements or parts; interpolated. In Japan, the basic codes are the Civil Code, the Commercial Code, the Penal Code, and procedural codes such as the Code of Criminal Procedure and the Code of Civil Procedure. —Roger E. Meiners, The Legal Environment of Business 3 Advanced VL Codes We start this chapter with codes for the integers. This is followed by many types of variable-length codes that are based on diverse principles, have been developed by different approaches, and have various useful properties and features. 3.1 VLCs for Integers Following Elias, it is customary to denote the standard binary representation of the integer n by ί(n). This representation can be considered a code (the beta code), but it does not satisfy the prefix property (because, for example, 2 = 102 is the prefix of 4 = 1002). The beta code has another disadvantage. Given a set of integers between 0 and n, we can represent each in 1+log2 n bits, a fixed-length representation. However, if the number of integers in the set is not known in advance (or if the largest integer is unknown), a fixed-length representation cannot be used and the natural solution is to assign variable-length codes to the integers. Any variable-length codes for integers should satisfy the following requirements: 1. Given an integer n, its codeword should be as short as possible and should be constructed from the magnitude, length, and bit pattern of n, without the need for any table lookups or other mappings. 2. Given a bitstream of variable-length codes, it should be easy to decode the next codeword and obtain an integer n even if n hasn’t been seen before. We will see that in many VLCs for integers, part of the binary representation of the integer is included in the codeword, and the rest of the codeword is side information indicating the length or precision of the encoded integer. Several codes for the integers are described in the first few sections of this chapter. Some can code only nonnegative integers and others can code only positive integers. A DOI 10.1007/978-1-84882-903-9_3, © Springer-Verlag London Limited 2010 D. Salomon, G. Motta, Handbook of Data Compression, 5th ed. 96 3. Advanced VL Codes VLC for positive integers can be extended to encode nonnegative integers by incrementing the integer before it is encoded and decrementing the result produced by decoding. A VLC for arbitrary integers can be obtained by a bijection, a mapping of the form 0 -1 1 -2 2 -3 3 -4 4 -5 5 · · · 1 2 3 4 5 6 7 8 9 10 11 · · · A function is bijective if it is one-to-one and onto. Perhaps the simplest variable-length code for integers is the well-known unary code. The unary code of the positive integer n is constructed as n - 1 bits of 1 followed by a single 0, or alternatively as n - 1 zeros followed by a single 1 (the three left columns of Table 3.1). The length of the unary code for the integer n is therefore n bits. The two rightmost columns of Table 3.1 show how the unary code can be extended to encode the nonnegative integers (which makes the codes one bit longer). The unary code is simple to construct and is useful in many applications, but it is not universal. Stone-age people indicated the integer n by marking n adjacent vertical bars on a stone, which is why the unary code is sometimes known as a stone-age binary and each of its n or (n-1) 1’s (or n or (n - 1) zeros) is termed a stone-age bit. Stone Age Binary? n Code Reverse Alt. code Alt. reverse 0 – – 0 1 1 0 1 10 01 2 10 01 110 001 3 110 001 1110 0001 4 1110 0001 11110 00001 5 11110 00001 111110 000001 Table 3.1: Some Unary Codes. It is easy to see that the unary code satisfies the prefix property, so it is instantaneous and can be used as a variable-length code. Since its length L satisfies L = n, we get 2-L = 2-n, so it makes sense to use this code in cases where the input data consists of integers n with exponential probabilities P(n) . 2-n. Given data that lends itself to the use of the unary code (i.e., a set of integers that satisfy P(n) . 2-n), we can assign unary codes to the integers and these codes will be as good as the Huffman codes with the advantage that the unary codes are trivial to encode and decode. In general, the unary code is used as part of other, more sophisticated, variable-length codes. Example: Table 3.2 lists the integers 1 through 6 with probabilities P(n) = 2-n, except that P(6) = 2-5 . 2-6. The table lists the unary codes and Huffman codes for the six integers, and it is obvious that these codes have the same lengths (except the code of 6, because this symbol does not have the correct probability). Every positive number was one of Ramanujan’s personal friends. —J. E. Littlewood 3.2 Start-Step-Stop Codes 97 n Prob. Unary Huffman 1 2-1 0 0 2 2-2 10 10 3 2-3 110 110 4 2-4 1110 1110 5 2-5 11110 11110 6 2-5 111110 11111 Table 3.2: Six Unary and Huffman Codes. 3.2 Start-Step-Stop Codes The unary code is ideal for compressing data that consists of integers n with probabilities P(n) . 2-n. If the data to be compressed consists of integers with different probabilities, it may benefit from one of the general unary codes (also known as start-step-stop codes). Such a code, proposed by [Fiala and Greene 89], depends on a triplet (start, step, stop) of nonnegative integer parameters. A set of such codes is constructed subset by subset as follows: 1. Set n = 0. 2. Set a = start + n Χ step. 3. Construct the subset of codes that start with n leading 1’s, followed by a single intercalary bit (separator) of 0, followed by a combination of a bits. There are 2a such codes. 4. Increment n by step. If n < stop, go to step 2. If n > stop, issue an error and stop. If n = stop, repeat steps 2 and 3 but without the single intercalary 0 bit of step 3, and stop. This construction makes it obvious that the three parameters have to be selected such that start + n Χ step will reach “stop” for some nonnegative n. The number of codes for a given triplet is normally finite and depends on the choice of parameters. It is given by 2stop+step - 2start 2step - 1 . Notice that this expression increases exponentially with parameter “stop,” so large sets of these codes can be generated even with small values of the three parameters. Notice that the case step = 0 results in a zero denominator and thus in an infinite set of codes. Tables 3.3 and 3.4 show the 680 codes of (3,2,9) and the 2044 codes of (2,1,10). Table 7.109 lists the number of codes of each of the general unary codes (2, 1, k) for k = 2, 3, . . . , 11. This table was calculated by the Mathematica command Table[2^(k+1)- 4,{k,2,11}]. Examples: 1. The triplet (n, 1, n) defines the standard (beta) n-bit binary codes, as can be verified by direct construction. The number of such codes is easily seen to be 2n+1 - 2n 21 - 1 = 2n. 98 3. Advanced VL Codes a = nth Number of Range of n 3 + n · 2 codeword codewords integers 0 3 0xxx 23 = 8 0–7 1 5 10xxxxx 25 = 32 8–39 2 7 110xxxxxxx 27 = 128 40–167 3 9 111xxxxxxxxx 29 = 512 168–679 Total 680 Table 3.3: The General Unary Code (3,2,9). a = nth Number of Range of n 2 + n · 1 codeword codewords integers 0 2 0xx 4 0–3 1 3 10xxx 8 4–11 2 4 110xxxx 16 12–27 3 5 1110xxxxx 32 28–59 · · · · · · · · · 8 10 11...1  8 xx...x    10 1024 1020–2043 Total 2044 Table 3.4: The General Unary Code (2,1,10). k : 2 3 4 5 6 7 8 9 10 11 (2, 1, k): 4 12 28 60 124 252 508 1020 2044 4092 Table 3.5: Number of General Unary Codes (2, 1, k) for k = 2, 3, . . . , 11. 2. The triplet (0, 0,.) defines the codes 0, 10, 110, 1110,. . . which are the unary codes but assigned to the integers 0, 1, 2,. . . instead of 1, 2, 3,. . . . 3. The triplet (0, 1,.) generates a variance of the Elias gamma code. 4. The triplet (k, k,.) generates another variance of the Elias gamma code. 5. The triplet (k, 0,.) generates the Rice code with parameter k. 6. The triplet (s, 1,.) generates the exponential Golomb codes of page 198. 7. The triplet (1, 1, 30) produces (230 - 21)/(21 - 1) . a billion codes. 8. Table 3.6 shows the general unary code for (10,2,14). There are only three code lengths since “start” and “stop” are so close, but there are many codes because “start” is large. The start-step-stop codes are very general. Often, a person gets an idea for a variable-length code only to find out that it is a special case of a start-step-stop code. Here is a typical example. We can design a variable-length code where each code consists of triplets. The first two bits of a triplet go through the four possible values 00, 01, 10, and 11, and the rightmost bit acts as a stop bit. If it is 1, we stop reading the code, otherwise we read another triplet of bits. Table 3.7 lists the first 22 codes. 3.3 Start/Stop Codes 99 a = nth Number of Range of n 10 + n · 2 codeword codewords integers 0 10 0x...x  10 210 = 1K 0–1023 1 12 10xx...x    12 212 = 4K 1024–5119 2 14 11xx...xx    14 214 = 16K 5120–21503 Total 21504 Table 3.6: The General Unary Code (10,2,14). n Codes n Codes 0 001 11 010 111 1 011 12 100 001 2 101 13 100 011 3 111 14 100 101 4 000 001 15 100 111 5 000 011 16 110 001 6 000 101 17 110 011 7 000 111 18 110 101 8 010 001 19 110 111 9 010 011 20 000 000 001 10 010 101 21 000 000 011 n Codes n Codes 0 1 00 11 01 01 11 1 1 01 12 01 10 00 2 1 10 13 01 10 01 3 1 11 14 01 10 10 4 01 00 00 15 01 10 11 5 01 00 01 16 01 11 00 6 01 00 10 17 01 11 01 7 01 00 11 18 01 11 10 8 01 01 00 19 01 11 11 9 01 01 01 20 001 00 00 00 10 01 01 10 21 001 00 00 01 Table 3.7: Triplet-Based Codes. Table 3.8: Same Codes Rearranged. To see why this is a start-step-stop code, we rewrite it by placing the stop bits on the left end, followed by pairs of code bits, as listed in Table 3.8. Examining these 22 codes shows that the code for the integer n starts with a unary code of j zeros followed by a single 1, followed by j + 1 pairs of code bits. There are four codes of length 3 (a 1-bit unary followed by a pair of code bits, corresponding to j = 0), 16 6-bit codes (a 2-bit unary followed by two pairs of code bits, corresponding to j = 1), 64 9-bit codes (corresponding to j = 2), and so on. A little tinkering shows that this is the start-step-stop code (2, 2,.). 3.3 Start/Stop Codes The start-step-stop codes are flexible. By carefully adjusting the values of the three parameters it is possible to construct sets of codes of many different lengths. However, the lengths of these codes are restricted to the values n+1+start+nΧstep (except for the last subset where the codes have lengths stop+start+stopΧstep). The start/stop codes of this section were conceived by Steven Pigeon and are described in [Pigeon 01a,b], where it is shown that they are universal. A set of these codes is fully specified by an 100 3. Advanced VL Codes array of nonnegative integer parameters (m0,m1, . . . , mt) and is constructed in subsets, similar to the start-step-stop codes, in the following steps: 1. Set i = 0 and a = m0. 2. Construct the subset of codes that start with i leading 1’s, followed by a single separator of 0, followed by a combination of a bits. There are 2a such codes. 3. Increment i by 1 and set a ‹ a +mi. 4. If i < t, go to step 2. Otherwise (i = t), repeat step 2 but without the single 0 intercalary, and stop. Thus, the parameter array (0, 3, 1, 2) results in the set of codes listed in Table 3.9. i a Codeword # of codes Length 0 2 0xx 4 3 1 5 10xxxxx 32 7 2 6 110xxxxxx 64 9 3 8 111xxxxxxxx 256 11 Table 3.9: The Start/Stop Code (2,3,1,2). The maximum code length is t+m0 +· · ·+mt and on average the start/stop code of an integer s is never longer than log2 s (the length of the binary representation of s). If an optimal set of such codes is constructed by an encoder and is used to compress a data file, then the only side information needed in the compressed file is the value of t and the array of t + 1 parameters mi (which are mostly small integers). This is considerably less than the size of a Huffman tree or the side information required by many other compression methods. The start/stop codes can also encode an indefinite number of arbitrarily large integers. Simply set all the parameters to the same value and set t to infinity. Steven Pigeon, the developer of these codes, shows that the parameters of the start/stop codes can be selected such that the resulting set of codes will have an average length shorter than what can be achieved with the start-step-stop codes for the same probability distribution. He also shows how the probability distribution can be employed to determine the best set of parameters for the code. In addition, the number of codes in the set can be selected as needed, in contrast to the start-step-stop codes that often result in more codes than needed. The Human Brain starts working the moment you are born and never stops until you stand up to speak in public! —George Jessel 3.4 Elias Codes 101 3.4 Elias Codes In his pioneering work [Elias 75], Peter Elias described three useful prefix codes. The main idea of these codes is to prefix the integer being encoded with an encoded representation of its order of magnitude. For example, for any positive integer n there is an integer M such that 2M . n < 2M+1. We can therefore write n = 2M + L where L is at most M bits long, and generate a code that consists of M and L. The problem is to determine the length of M and this is solved in different ways by the various Elias codes. Elias denoted the unary code of n by .(n) and the standard binary representation of n, from its most-significant 1, by ί(n). His first code was therefore designated . (gamma). Elias gamma code .(n) for positive integers n is simple to encode and decode and is also universal. Encoding. Given a positive integer n, perform the following steps: 1. Denote by M the length of the binary representation ί(n) of n. 2. Prepend M - 1 zeros to it (i.e., the .(n) code without its terminating 1). Step 2 amounts to prepending the length of the code to the code, in order to ensure unique decodability. The length M of the integer n is, from Equation (2.1), 1 + log2 n, so the length of .(n) is 2M - 1 = 2log2 n + 1. (3.1) We later show that this code is ideal for applications where the probability of n is 1/(2n2). An alternative construction of the gamma code is as follows: 1. Find the largest integer N such that 2N . n < 2N+1 and write n = 2N + L. Notice that L is at most an N-bit integer. 2. Encode N in unary either as N zeros followed by a 1 or N 1’s followed by a 0. 3. Append L as an N-bit number to this representation of N. Section 3.8 describes yet another way to construct the same code. Section 3.15 shows a connection between this code and certain binary search trees. Peter Elias 1 = 20 +0 = 1 10 = 23 + 2 = 0001010 2 = 21 + 0 = 010 11 = 23 + 3 = 0001011 3 = 21 + 1 = 011 12 = 23 + 4 = 0001100 4 = 22 + 0 = 00100 13 = 23 + 5 = 0001101 5 = 22 + 1 = 00101 14 = 23 + 6 = 0001110 6 = 22 + 2 = 00110 15 = 23 + 7 = 0001111 7 = 22 + 3 = 00111 16 = 24 + 0 = 000010000 8 = 23 + 0 = 0001000 17 = 24 + 1 = 000010001 9 = 23 + 1 = 0001001 18 = 24 + 2 = 000010010 Table 3.10: 18 Elias Gamma Codes. Table 3.10 lists the first 18 gamma codes, where the L part is in italics (see also Table 10.29 and the C1 code of Table 3.20). 102 3. Advanced VL Codes In his 1975 paper, Elias describes two versions of the gamma code. The first version (titled .) is encoded as follows: 1. Generate the binary representation ί(n) of n. 2. Denote the length |ί(n)| of ί(n) by M. 3. Generate the unary u(M) representation of M as M - 1 zeros followed by a 1. 4. Follow each bit of ί(n) by a bit of u(M). 5. Drop the leftmost bit (the leftmost bit of ί(n) is always 1). Thus, for n = 13 we prepare ί(13) =―1 ―1 ―0 ―1, so M = 4 and u(4) = 0001, resulting in ―10―10―00―11. The final code is .(13) = 0―10―00―11. In general, the length of the gamma code for the integer i is 1 + 2log2 i. The second version, dubbed ., moves the bits of u(M) to the left. Thus .(13) = 0001|―1 ―0 ―1. The gamma codes of Table 3.10 are Elias’s . codes. Both gamma versions are universal. Decoding is also simple and is done in two steps: 1. Read zeros from the code until a 1 is encountered. Denote the number of zeros by N. 2. Read the next N bits as an integer L. Compute n = 2N + L. It is easy to see that this code can be used to encode positive integers even in cases where the largest integer is not known in advance. Also, this code grows slowly (see Figure 3.38), so it is a good candidate for compressing integer data where small integers are common and large ones are rare. It is easy to understand why the gamma code (and in general, all variable-length codes) should be used only when it is known or estimated that the distribution of the integers to be encoded is close to the ideal distribution for the given code. The simple computation here assumes that the first n integers are given and are distributed uniformly (i.e., each appears with the same probability). We compute the average gamma code length and show that it is much larger than the length of the fixed-size binary codes of the same integers. The length of the Elias gamma code for the integer i is 1 + 2log2 i. Thus, the average of the gamma codes of the first n integers is a = n + 2n i=1log2 i n . The length of the fixed-size codes of the same integers is b = log2 n. Figure 3.11 is a plot of the ratio a/b and it is easy to see that this value is greater than 1, indicating that for flat distributions of the integers the fixed-length (beta) code is better than the gamma code. The curve in the figure generally goes up, but features sudden drops (where it remains greater than 1) at points where n = 2k + 1 for integer k. Elias delta code. In his gamma code, Elias prepends the length of the code in unary (.). In his next code, . (delta), he prepends the length in binary (ί). Thus, the Elias delta code, also for the positive integers, is slightly more complex to construct. Encoding a positive integer n, is done in the following steps: 1. Write n in binary. The leftmost (most-significant) bit will be a 1. 3.4 Elias Codes 103         gamma[i_] := 1. + 2 Floor[Log[2, i]]; Plot[Sum[gamma[j], {j,1,n}]/(n Ceiling[Log[2,n]]), {n,1,200}] Figure 3.11: Gamma Code Versus Binary Code. 2. Count the bits, remove the leftmost bit of n, and prepend the count, in binary, to what is left of n after its leftmost bit has been removed. 3. Subtract 1 from the count of step 2 and prepend that number of zeros to the code. When these steps are applied to the integer 17, the results are: 17 = 100012 (five bits). Remove the leftmost 1 and prepend 5 = 1012 yields 101|0001. Three bits were added, so we prepend two zeros to obtain the delta code 00|101|0001. To compute the length of the delta code of n, we notice that step 1 generates (from Equation (2.1)) M = 1+log2 n bits. For simplicity, we omit the  and  and observe that M = 1 + log2 n = log2 2 + log2 n = log2(2n). The count of step 2 isM, whose length C is therefore C = 1+log2M = 1+log2(log2(2n)) bits. Step 2 therefore prepends C bits and removes the leftmost bit of n. Step 3 prepends C - 1 = log2M = log2(log2(2n)) zeros. The total length of the delta code is therefore the 3-part sum log2(2n) + [1 + log2 log2(2n)] - 1 + log2 log2(2n) = 1+log2 n + 2log2 log2(2n). step 1  step 2   step 3  (3.2) Figure 3.38 illustrates the length graphically. We show below that this code is ideal for data where the integer n occurs with probability 1/[2n(log2(2n))2]. An equivalent way to construct the delta code employs the gamma code: 1. Find the largest integer N such that 2N . n < 2N+1 and write n = 2N + L. Notice that L is at most an N-bit integer. 2. Encode N + 1 with the Elias gamma code. 3. Append the binary value of L, as an N-bit integer, to the result of step 2. 104 3. Advanced VL Codes When these steps are applied to n = 17, the results are: 17 = 2N +L = 24+1. The gamma code of N + 1 = 5 is 00101, and appending L = 0001 to this yields 00101|0001. Table 3.12 lists the first 18 delta codes, where the L part is in italics. See also the related code C3 of Table 3.20, which has the same length. 1 = 20 + 0 › |L| = 0 ›1 10=23 + 2 › |L| = 3 › 00100010 2 = 21 + 0 › |L| = 1 › 0100 11 = 23 + 3 › |L| = 3 › 00100011 3 = 21 + 1 › |L| = 1 › 0101 12 = 23 + 4 › |L| = 3 › 00100100 4 = 22 + 0 › |L| = 2 › 01100 13 = 23 + 5 › |L| = 3 › 00100101 5 = 22 + 1 › |L| = 2 › 01101 14 = 23 + 6 › |L| = 3 › 00100110 6 = 22 + 2 › |L| = 2 › 01110 15 = 23 + 7 › |L| = 3 › 00100111 7 = 22 + 3 › |L| = 2 › 01111 16 = 24 + 0 › |L| = 4 › 001010000 8 = 23 + 0 › |L| = 3 › 00100000 17 = 24 + 1 › |L| = 4 › 001010001 9 = 23 + 1 › |L| = 3 › 00100001 18 = 24 + 2 › |L| = 4 › 001010010 Table 3.12: 18 Elias Delta Codes. Decoding is done in the following steps: 1. Read bits from the code until you can decode an Elias gamma code. Call the decoded result M + 1. This is done in the following substeps: 1.1 Count the leading zeros of the code and denote the count by C. 1.2 Examine the leftmost 2C + 1 bits (C zeros, followed by a single 1, followed by C more bits). This is the decoded gamma code M + 1. 2. Read the next M bits. Call this number L. 3. The decoded integer is 2M + L. In the case of n = 17, the delta code is 001010001. We skip two zeros, so C = 2. The value of the leftmost 2C +1 = 5 bits is 00101 = 5, so M +1 = 5. We read the next M = 4 bits 0001, and end up with the decoded value 2M + L = 24 + 1 = 17. To better understand the application and performance of these codes, we need to identify the type of data it compresses best. Given a set of symbols ai, where each symbol occurs in the data with probability Pi and the length of its code is li bits, the average code length is the sum Pili and the entropy (the smallest number of bits required to represent the symbols) is [-Pi log2 Pi]. The redundancy (Equation (2.2)) is the difference i Pili -i[-Pi log2 Pi] and we are looking for probabilites Pi that will minimize this difference. In the case of the gamma code, li = 1+2log2 i. If we select symbol probabilities Pi = 1/(2i2) (a power law distribution of probabilities, where the first 10 values are 0.5, 0.125, 0.0556, 0.03125, 0.02, 0.01389, 0.0102, 0.0078, 0.00617, and 0.005, see also Table 3.21), both the average code length and the entropy become the identical sums i 1 + 2 log i 2i2 , 3.4 Elias Codes 105 indicating that the gamma code is asymptotically optimal for this type of data. A power law distribution of values is dominated by just a few symbols and especially by the first. Such a distribution is very skewed and is therefore handled very well by the gamma code which starts very short. In an exponential distribution, in contrast, the small values have similar probabilities, which is why data with this type of statistical distribution is compressed better by a Rice code (Section 3.25). In the case of the delta code, li = 1 + log i + 2 log log(2i). If we select symbol probabilities Pi = 1/[2i(log(2i))2] (where the first five values are 0.5, 0.0625, 0.025, 0.0139, and 0.009), both the average code length and the entropy become the identical sums i log 2 + log i + 2 log log(2i) 2i(log(2i))2 , indicating that the redundancy is zero and the delta code is therefore asymptotically optimal for this type of data. Section 3.15 shows a connection between a variant of the delta code and certain binary search trees. The phrase “working mother” is redundant. —Jane Sellman The Elias omega code. Unlike the previous Elias codes, the omega code uses itself recursively to encode the prefix M, which is why it is sometimes referred to as a recursive Elias code. The main idea is to prepend the length of n to n as a group of bits that starts with a 1, then prepend the length of the length, as another group, to the result, and continue prepending lengths until the last length is 2 or 3 (and therefore fits in two bits). In order to distinguish between a length group and the last, rightmost group (of n itself), the latter is followed by a delimiter of 0, while each length group starts with a 1. Encoding a positive integer n is done recursively in the following steps: 1. Initialize the code-so-far to 0. 2. If the number to be encoded is 1, stop; otherwise, prepend the binary representation of n to the code-so-far. Assume that we have prepended L bits. 3. Repeat step 2, with the binary representation of L - 1 instead of n. The integer 17 is therefore encoded by (1) a single 0, (2) prepended by the 5-bit binary value 10001, (3) prepended by the 3-bit value of 5-1 = 1002, and (4) prepended by the 2-bit value of 3 - 1 = 102. The result is 10|100|10001|0. Table 3.13 lists the first 18 omega codes (see also Table 3.17). Note that n = 1 is handled as a special case. Decoding is done in several nonrecursive steps where each step reads a group of bits from the code. A group that starts with a zero signals the end of decoding. 1. Initialize n to 1. 2. Read the next bit. If it is 0, stop. Otherwise read n more bits, assign the group of n + 1 bits to n, and repeat this step. 106 3. Advanced VL Codes 1 0 10 11 1010 0 2 10 0 11 11 1011 0 3 11 0 12 11 1100 0 4 10 100 0 13 11 1101 0 5 10 101 0 14 11 1110 0 6 10 110 0 15 11 1111 0 7 10 111 0 16 10 100 10000 0 8 11 1000 0 17 10 100 10001 0 9 11 1001 0 18 10 100 10010 0 Table 3.13: 18 Elias Omega Codes. Some readers may find it easier to understand these steps rephrased as follows. 1. Read the first group, which will either be a single 0, or a 1 followed by n more digits. If the group is a 0, the value of the integer is 1; if the group starts with a 1, then n becomes the value of the group interpreted as a binary number. 2. Read each successive group; it will either be a single 0, or a 1 followed by n more digits. If the group is a 0, the value of the integer is n; if it starts with a 1, then n becomes the value of the group interpreted as a binary number. Example. Decode 10|100|10001|0. The decoder initializes n = 1 and reads the first bit. It is a 1, so it reads n = 1 more bit (0) and assigns n = 102 = 2. It reads the next bit. It is a 1, so it reads n = 2 more bits (00) and assigns the group 100 to n. It reads the next bit. It is a 1, so it reads four more bits (0001) and assigns the group 10001 to n. The next bit read is 0, indicating the end of decoding. The omega code is constructed recursively, which is why its length |.(n)| can also be computed recursively. We define the quantity lk(n) recursively by l1(n) = log2 n and li+1(n) = l1(li(n)). Equation (2.1) tells us that |ί(n)| = l1(n) + 1 (where ί is the standard binary representation), and this implies that the length of the omega code is given by the sum |.(n)| = k i=1 ί(lk-i(n)) + 1 = 1+ k i=1 (li(n) + 1), where the sum stops at the k that satisfies lk(n) = 1. From this, Elias concludes that the length satisfies |.(n)| . 1 + 5 2 log2 n. A quick glance at any table of these codes shows that their lengths fluctuate. In general, the length increases slowly as n increases, but when a new length group is added, which happens when n = 22k for any positive integer k, the length of the code increases suddenly by several bits. For k values of 1, 2, 3, and 4, this happens when n reaches 4, 16, 256, and 65,536. Because the groups of lengths are of the form “length,” “log(length),” “log(log(length)),” and so on, the omega code is sometimes referred to as a logarithmic-ramp code. Table 3.14 compares the length of the gamma, delta, and omega codes. It shows that the delta code is asymptotically best, but if the data consists mostly of small numbers 3.5 RBUC, Recursive Bottom-Up Coding 107 Values Gamma Delta Omega 1 1 1 2 2 3 4 3 3 3 4 4 4 5 5 4 5–7 5 5 5 8–15 7 8 6–7 16–31 9 9 7–8 32–63 11 10 8–10 64–88 13 11 10 100 13 11 11 1000 19 16 16 104 27 20 20 105 33 25 25 105 39 28 30 Table 3.14: Lengths of Three Elias Codes. (less than 8) and there are only a few large integers, then the gamma code performs better. Beware of bugs in the above code; I have only proved it correct, not tried it. —Donald Knuth 3.5 RBUC, Recursive Bottom-Up Coding The RBUC (recursive bottom-up coding) method discussed here [Moffat and Anh 06] is based on the notion of a selector. We know from Equation (2.1) that the length b of the integer n is 1+log2 n = log2(1+n). Thus, given a set S of nonnegative integers, we can use the standard binary (ί) code to encode each integer in b = log2(1 + x) bits, where x is the largest integer in the set. We can consider b the selector for S, and since the same b is used for the entire set, it can be termed a global selector. This type of encoding is simple but extremely ineffectual because the codes of many integers smaller than x will have redundant leading zeros. In the other extreme we find the Elias Gamma code (Section 3.4). Given a positive integer n, its gamma code is constructed in the following steps: 1. Denote by b the length of the binary representation ί(n) of n. 2. Prepend b - 1 zeros to ί(n). Thus, each gamma code includes its own selector b. This code is more efficient than ί but takes more steps to construct. Once we look at the ί and . codes as extreme approaches to selectors, it is natural to search for a more flexible approach, and RBUC is an important step in this direction. It represents a compromise between the ί and . codes and it results in good compression when applied to the right type of data. 108 3. Advanced VL Codes Given a set of m integers, the RBUC encoder partitions it into segments of size s (except the last segment, which may be shorter), finds the largest integer in each segment i, uses it to determine a selector bi for the segment, and employs the standard ί code to encode the elements of the segment in bi bits each. It is obvious that RBUC can yield good compression if each segment is uniform or close to uniform, and one example of such data is inverted indexes [Witten et al. 99]. An inverted index is a data structure that stores a mapping from an index item (normally a word) to its locations in a set of documents. Thus, the inverted index for index-item f may be the pair [f, (2, 5, 6, 8, 11, 12, 13, 18)] where the integers 2, 5, 6, and so on are document numbers. It is known experimentally that such lists of document numbers exhibit local uniformity and are therefore natural candidates for RBUC encoding. Even with ideal data, there is a price to pay for the increased performance of RBUC. Compressing each segment with a separate selector bi must be followed by recursively compressing the set of selectors. There are m/s selectors in this set and they are encoded by partitioning the set into subsets of size s each, determining a selector ci for each, and encoding with ί. The set of selectors ci has to be encoded too, and this process continues recursively log2m times until a set with just one selector is obtained. This selector is encoded with an Elias code, and the encoder also has to encode the value of m, as overhead, for the decoder’s use. Here is an example. Given the m = 13 set of integers (15, 6, 2, 3, 0, 0, 0, 0, 4, 5, 1, 7, 8) we select s = 2 and partition the set into seven subsets of two integers each (except the last subset). Examining the elements of each subset, it is easy to determine the set of selectors bi = (4, 3, 0, 0, 3, 3, 4) and encode the 13 integers. Notice that the four zeros are encoded into zero bits, thereby contributing to compression effectiveness. The process iterates four times (notice that log2m = 3.7) and is summarized in Table 3.15. The last selector, 2, is encoded as the gamma code 010, and the encoder also has to encode m = 13, perhaps also as a gamma code. The codes are then collected from the top of the table and are concatenated to become the compressed stream. Notice that the original message and all the iterations must be computed and kept in memory until the moment of output, which is the main drawback of this method. We can say that RBUC is an offline algorithm. 010 ei = (2) 10|10 di = (2, 2) 11|00|01|11 ci = (3, 0, 2, 2) 100|011|011|011|100 bi = (4, 3, 0, 0, 3, 3, 4) 1111|0110|010|011|100|101|001|111|1000 15, 6, 2, 3, 0, 0, 0, 0, 4, 5, 1, 7, 8 Table 3.15: An RBUC Example with m = 13 and s = 2. Assuming that the largest integer in set i is a, the selector bi for the set will be log2(1+a). Thus, the set of selectors (bi) consists of integers that are the logarithms of the original data symbols and are therefore much smaller. The set of selectors (ci) will be much smaller still, and so on. Since all the data symbols and selectors are nonnegative 3.5 RBUC, Recursive Bottom-Up Coding 109 integers, consecutive sets of selectors become logarithmically more uniform and reduce any nonhomogeneity in the original data. The need to iterate log2m times and compress sets of selectors reduces the effectiveness and increases the execution time of RBUC. However, the sets of selectors become more and more uniform, which suggests the use of larger sets (i.e., larger values of s) in each iteration and thus fewer iterations. If we double s in each iteration, the number of iterations reduces to log2m, and if we square s each time, the number of iterations shrinks to the much smaller log2 log2m. Another minor improvement can be incorporated when we observe that the set of data items 1, 7 are encoded in three bits to become 001 and 111 (Table 3.15), but can also be encoded to 001 and 11. Once we see the 3-bit codeword 001, we know that the other codeword in this set must start with a 1, so this 1 can be omitted. In general, if s-1 codewords of a set start with a zero, then the remaining codeword must start with a 1 which can be omitted. 3.5.1 BASC, Binary Adaptive Sequential Coding RBUC is an offline algorithm, which is why its developers also came up with a similar encoding method that is both online and adaptive. (“Online” means that the codewords can be emitted as soon as they are computed and there is no need to save large sets of numbers in memory.) They termed it binary adaptive sequential coding (BASC). The principle of BASC is to determine a selector b for each data symbol based on the selector of the preceding symbol. Once b has been determined, the current symbol is encoded, the codeword is emitted, and only b need be saved, for use with the next data symbol. The principle of BASC is to determine a selector b for each data symbol based on the selector b of the preceding symbol. Once b has been determined, the current symbol is encoded, the codeword is emitted, and b is set to b. Only b need be saved, for use with the next data symbol. Specifically, the BASC algorithm operates as follows: The initial value of b is either input by the user or is determined from the first data symbol. Given the current data symbol xi, its length b = log2(1 + xi) is computed. If b . b, then a zero is emitted, followed by the ί code of xi as a b-bit number. Otherwise, the encoder emits (b - b) 1’s, followed by a single zero, followed by the least-significant b -1 bits of xi (the mostsignificant bit of xi must be a 1, and so can be ignored). This process is illustrated by the 13-symbol message (15, 6, 2, 3, 0, 0, 0, 0, 4, 5, 1, 7, 8). Assume that the initial b is 5. The first symbol, 15, yields b = 4, which is less than b. The encoder generates and outputs a zero followed by 15 as a b-bit number 0|01111 and b is set to 4. The second symbol, 6, yields b = 3, so 0| 0110 is emitted and b is set to 3. The third symbol, 2, yields b = 2, so 0| 010 is output and b is set to 2. The next symbol, 3, results in b = 2 and 0|11 is output, followed by b ‹ 2. The first of the four zeros yields b = 0. The codeword 0|00 is emitted and b is set to 0. Each successive zero produces b = 0 and is output as 0. The next symbol, 4, yields b = 3 which is greater than b. The encoder emits b -b = 3-0 = 3 1’s, followed by a zero, followed by the two least-significant bits 00 of the 4, thus 111|0|00. The current selector b is set to 3 and the rest of the example is left as an easy exercise. 110 3. Advanced VL Codes The decoder receives the values of m (13) and the initial b (5) as overhead, so it knows how many codewords to read and how to read the first one. If a codeword starts with a zero, the decoder reads b bits, converts them to an integer n, computes the number of bits required for n, and sets b to that number. If a codeword starts with a 1, the decoder reads consecutive 1s until the first zero, skips the zero, and stores the number of 1s in c. This number is b - b. The decoder has b from the previous step, so it can compute b = c + b and read b - 1 bits from the current codeword. The decoder prepends a 1 to those bits, outputs the result as the next symbol, and sets b to b. The codewords produced by BASC are similar to the Elias gamma codes for the integers, the main difference is the adaptive step where b is set to b. The developers of this method propose two variations to this simple step. The first variation is to set b to the average of several previous bs. This can be thought of as smoothing any large differences in consecutive bs because of uneveness in the input data. The second, damping, variation tries to achieve the same goal by limiting the changes in b to one unit at a time. The developers of RBUC and BASC list results of tests comparing the performance of these algorithms to the Elias gamma codes, Huffman codes, and the interpolative coding algorithm of Section 3.28. The results indicate that BASC and RBUC offer new tradeoffs between compression effectiveness and decoding and are therefore serious competitors to the older, well-known, and established compression algorithms. 3.6 Levenstein Code This little-known code for the nonnegative integers was conceived in 1968 by Vladimir Levenshtein [Levenstein 06]. Both encoding and decoding are multistep processes. Encoding. The Levenstein code of zero is a single 0. To code a positive number n, perform the following: 1. Set the count variable C to 1. Initialize the code-so-far to the empty string. 2. Take the binary value of n without its leading 1 and prepend it to the code-so-far. 3. Let M be the number of bits prepended in step 2. 4. If M = 0, increment C by 1, go to and execute step 2 with M instead of n. 5. If M = 0, prepend C 1’s followed by a 0 to the code-so-far and stop. n Levenstein code n Levenstein code 0 0 9 1110 1 001 1 10 10 1110 1 010 2 110 0 11 1110 1 011 3 110 1 12 1110 1 100 4 1110 0 00 13 1110 1 101 5 1110 0 01 14 1110 1 110 6 1110 0 10 15 1110 1 111 7 1110 0 11 16 11110 0 00 0000 8 1110 1 000 17 11110 0 00 0001 Table 3.16: 18 Levenstein Codes. 3.7 Even–Rodeh Code 111 Table 3.16 lists some of these codes. Spaces have been inserted to indicate the individual parts of each code. As an exercise, the reader may verify that the Levenstein codes for 18 and 19 are 11110|0|00|0010 and 11110|0|00|0011, respectively. Decoding is done as follows: 1. Set count C to the number of consecutive 1’s preceding the first 0. 2. If C = 0, the decoded value is zero, stop. 3. Set N = 1, and repeat step 4 (C - 1) times. 4. Read N bits, prepend a 1, and assign the resulting bitstring to N (thereby erasing the previous value of N). The string assigned to N in the last iteration is the decoded value. The Levenstein code of the positive integer n is always one bit longer than the Elias omega code of n. 3.7 Even–Rodeh Code The principle behind the Elias omega code occurred independently to S. Even and M. Rodeh [Even and Rodeh 78]. Their code is similar to the omega code, the main difference being that lengths are prepended until a 3-bit length is reached and becomes the leftmost group of the code. For example, the Even–Rodeh code of 2761 is 100|1100|101011001001|0 (4, 12, 2761, and 0). The authors prove, by induction on n, that every nonnegative integer can be encoded in this way. They also show that the length l(n) of their code satisfies l(n) . 4+2L(n) where L(n) is the length of the binary representation (beta code) of n, Equation (2.1). The developers show how this code can be used as a comma to separate variablelength symbols in a string. Given a string of symbols ai, precede each ai with the Even–Rodeh code R of its length li. The codes and symbols can then be concatenated into a single string R(l1)a1R(l2)a2 . . . R(lm)am000 that can be uniquely separated into the individual symbols because the codes act as separators (commas). Three zeros act as the string delimiter. The developers also prove that the extra length of the codes shrinks asymptotically to zero as the string becomes longer. For strings of length 210 bits, the overhead is less than 2%. Table 3.17 (after [Fenwick 96a]) lists several omega and Even–Rodeh codes. Table 3.18 (after [Fenwick 96a]) illustrates the different points where the lengths of these codes increase. The length of the omega code increases when n changes from the form 2m - 1 (where the code is shortest relative to n) to 2m (where it is longest). The Even–Rodeh code behaves similarly, but may be slightly shorter or longer for various intervals of n. 112 3. Advanced VL Codes n Omega Even–Rodeh 0 — 000 1 0 001 2 10 0 010 3 11 0 011 4 10 100 0 100 0 7 10 111 0 111 0 8 11 1000 0 100 1000 0 15 11 1111 0 100 1111 0 16 10 100 10000 0 101 10000 0 32 10 101 100000 0 110 100000 0 100 10 110 1100100 0 111 1100100 0 1000 10 110 1100100 0 110 1100100 0 Table 3.17: Several Omega and Even–Rodeh Codes. Values . ER 1 1 3 2–3 3 3 4–7 6 4 8–15 7 8 16–31 11 9 32–63 12 10 64–127 13 11 128–255 14 17 256–512 21 18 Table 3.18: Different Lengths. 3.8 Punctured Elias Codes The punctured Elias codes for the integers were designed by Peter Fenwick in an attempt to improve the performance of the Burrows–Wheeler transform [Burrows and Wheeler 94]. The codes are described in the excellent, 15-page technical report [Fenwick 96a]. The term “punctured” comes from the field of error-control codes. Often, a codeword for error-detection or error-correction consists of the original data bits plus a number of check bits. If some check bits are removed, to shorten the codeword, the resulting code is referred to as punctured. We start with the Elias gamma code. Section 3.4 describes two ways to construct this code, and here we consider a third approach to the same construction. Write the binary value of n, reverse its bits so its rightmost bit is now a 1, and prepend flags for the bits of n. For each bit of n, create a flag of 0, except for the last (rightmost) bit (which is a 1), whose flag is 1. Prepend the flags to the reversed n and remove the rightmost bit. Thus, 13 = 11012 is reversed to yield 1011. The four flags 0001 are prepended and the rightmost bit of 1011 is removed to yield the final gamma code 0001|101. The punctured code eliminates the flags for zeros and is constructed as follows. Write the binary value of n, reverse its bits, and prepend flags to indicate the number of 1’s in n. For each bit of 1 in n we prepare a flag of 1, and terminate the flags with a single 0. Thus, 13 = 11012 is reversed to 1011. It has three 1’s, so the flags 1110 are prepended to yield 1110|1011. We call this punctured code P1 and notice that it starts with a 1 (there is at least one flag, except for the P1 code of n = 0) and also ends with a 1 (because the original n, whose MSB is a 1, has been reversed). We can therefore construct another punctured code P2 such that P2(n) equals P1(n + 1) with its most-significant 1 removed. Table 3.19 lists examples of P1 and P2. The feature that strikes us most is that the codes are generally getting longer as n increases, but are also getting shorter from time to time, for n values that are 1 less than a power of 2. For small values of n, these codes are often a shade longer than the gamma code, but for large values they average about 1.5 log2 n bits, shorter than the 2log2 n + 1 bits of the gamma code. 3.9 Other Prefix Codes 113 One design consideration for these codes was their expected behavior when applied to data with a skewed distribution and more smaller values than larger values. Smaller binary integers tend to have fewer 1’s, so it was hoped that the punctures would reduce the average code length below 1.5 log2 n bits. Later work by Fenwick showed that this hope did not materialize. n P1 P2 n P1 P2 0 0 01 11 11101101 100011 1 101 001 12 1100011 1101011 2 1001 1011 13 11101011 1100111 3 11011 0001 14 11100111 11101111 4 10001 10101 15 111101111 000001 5 110101 10011 16 1000001 1010001 6 110011 110111 . . . 7 1110111 00001 31 11111011111 0000001 8 100001 101001 32 10000001 10100001 9 1101001 100101 33 110100001 10010001 10 1100101 1101101 Table 3.19: Two Punctured Elias Codes. And as they carried him away Our punctured hero was heard to say, When in this war you venture out, Best never do it dressed as a Kraut! —Stephen Ambrose, Band of Brothers 3.9 Other Prefix Codes Four prefix codes, C1 through C4, are presented in this section. We denote by B(n) the binary representation of the integer n (the beta code). Thus, |B(n)| is the length, in bits, of this representation. We also use B(n) to denote B(n) without its most significant bit (which is always 1). Code C1 consists of two parts. To code the positive integer n, we first generate the unary code of |B(n)| (the size of the binary representation of n), then append B(n) to it. An example is n = 16 = 100002. The size of B(16) is 5, so we start with the unary code 11110 (or 00001) and append B(16) = 0000. Thus, the complete code is 11110|0000 (or 00001|0000). Another example is n = 5 = 1012 whose code is 110|01. The length of C1(n) is 2log2 n + 1 bits. Notice that this code is identical to the general unary code (0, 1,.) and is closely related to the Elias gamma code. Code C2 is a rearrangement of C1 where each of the 1 + log2 n bits of the first part (the unary code) of C1 is followed by one of the bits of the second part. Thus, code C2(16) = 101010100 and C2(5) = 10110. 114 3. Advanced VL Codes Code C3 starts with |B(n)| coded in C2, followed by B(n). Thus, 16 is coded as C2(5) = 10110 followed by B(16) = 0000, and 5 is coded as code C2(3) = 110 followed by B(5) = 01. The length of C3(n) is 1+log2 n + 2log2(1 + log2 n) (same as the length of the Elias delta code, Equation (3.2)). Code C4 consists of several parts. We start with B(n). To its left we prepend the binary representation of |B(n)| - 1 (one less than the length of n). This continues recursively, until a 2-bit number is written. A 0 is then appended to the right of the entire code, to make it uniquely decodable. To encode 16, we start with 10000, prepend |B(16)|-1 = 4 = 1002 to the left, then prepend |B(4)|-1 = 2 = 102 to the left of that, and finally append a 0 on the right. The result is 10|100|10000|0. To encode 5, we start with 101, prepend |B(5)| - 1 = 2 = 102 to the left, and append a 0 on the right. The result is 10|101|0. Comparing with Table 3.13 shows that C4 is the omega code. (The 0 on the right make the code uniquely decodable because each part of C4 is the standard binary code of some integer, so it starts with a 1. A start bit of 0 is therefore a signal to the decoder that this is the last part of the code.) Table 3.20 shows examples of the four codes above, as well as B(n) and B(n). The lengths of the four codes shown in the table increase as log2 n, in contrast with the length of the unary code, which increases as n. These codes are therefore good choices in cases where the data consists of integers n with probabilities that satisfy certain conditions. Specifically, the length L of the unary code of n is L = n = log2 2n, so it is ideal for the case where P(n) = 2-L = 2-n. The length of code C1(n) is L = 1+2log2 n = log2 2 + log2 n2 = log2(2n2), so it is ideal for the case where P(n) = 2-L = 1 2n2 . The length of code C3(n) is L = 1+log2 n + 2log2(1 + log2 n) = log2 2 + 2log log2 2n + log2 n, so it is ideal for the case where P(n) = 2-L = 1 2n(log2 n)2 . Table 3.21 shows the ideal probabilities that the first eight positive integers should have for the unary, C1, and C3 codes to be used. More prefix codes for the positive integers, appropriate for special applications, may be designed by the following general approach. Select positive integers vi and combine them in a list V (which may be finite or infinite according to needs). The code of the positive integer n is prepared in the following steps: 1. Find k such that k-1 i=1 vi < n . k i=1 vi. 2. Compute the difference d = n - k-1 i=1 vi- 1. 3.9 Other Prefix Codes 115 n B(n) B(n) C1 C2 C3 C4 1 1 0| 0 0| 0 2 10 0 10|0 100 100|0 10|0 3 11 1 10|1 110 100|1 11|0 4 100 00 110|00 10100 110|00 10|100|0 5 101 01 110|01 10110 110|01 10|101|0 6 110 10 110|10 11100 110|10 10|110|0 7 111 11 110|11 11110 110|11 10|111|0 8 1000 000 1110|000 1010100 10100|000 11|1000|0 9 1001 001 1110|001 1010110 10100|001 11|1001|0 10 1010 010 1110|010 1011100 10100|010 11|1010|0 11 1011 011 1110|011 1011110 10100|011 11|1011|0 12 1100 100 1110|100 1110100 10100|100 11|1100|0 13 1101 101 1110|101 1110110 10100|101 11|1101|0 14 1110 110 1110|110 1111100 10100|110 11|1110|0 15 1111 111 1110|111 1111110 10100|111 11|1111|0 16 10000 0000 11110|0000 101010100 10110|0000 10|100|10000|0 31 11111 1111 11110|1111 111111110 10110|1111 10|100|11111|0 32 100000 00000 111110|00000 10101010100 11100|00000 10|101|100000|0 63 111111 11111 111110|11111 11111111110 11100|11111 10|101|111111|0 64 1000000 000000 1111110|000000 1010101010100 11110|000000 10|110|1000000|0 127 1111111 111111 1111110|111111 1111111111110 11110|111111 10|110|1111111|0 128 10000000 0000000 11111110|0000000 101010101010100 1010100|0000000 10|111|10000000|0 255 11111111 1111111 11111110|1111111 111111111111110 1010100|1111111 10|111|11111111|0 Table 3.20: Some Prefix Codes. n Unary C1 C3 1 0.5 0.5000000 2 0.25 0.1250000 0.2500000 3 0.125 0.0555556 0.0663454 4 0.0625 0.0312500 0.0312500 5 0.03125 0.0200000 0.0185482 6 0.015625 0.0138889 0.0124713 7 0.0078125 0.0102041 0.0090631 8 0.00390625 0.0078125 0.0069444 Table 3.21: Ideal Probabilities of Eight Integers for Three Codes. 116 3. Advanced VL Codes The largest value of n isk 1 vi, so the largest value of d isk i vi-k-1 1 vi-1 = vk-1, a number that can be written in log2 vk bits. The number d is encoded, using the standard binary code, either in this number of bits, or if d < 2log2 vk-vk, it is encoded in log2 vk bits. 3. Encode n in two parts. Start with k encoded in some prefix code, and concatenate the binary code of d. Since k is coded in a prefix code, any decoder would know how many bits to read for k. After reading and decoding k, the decoder can compute the value 2log2 vk - vk which tells it how many bits to read for d. A simple example is the infinite sequence V = (1, 2, 4, 8, . . . , 2i-1, . . .) with k coded in unary. The integer n = 10 satisfies 3 i=1 vi < 10 . 4 i=1 vi, so k = 4 (with unary code 1110) and d = 10 - 3 i=1 vi- 1 = 2. Our values vi are powers of 2, so log2 vi is an integer and 2log2 vk equals vi. Thus, the length of d in our example is log2 vi = log2 8 = 3 and the code of 10 is 1110|010. 3.10 Ternary Comma Code Binary (base 2) numbers are based on the two bits 0 and 1. Similarly, ternary (base 3) numbers are based on the three digits (trits) 0, 1, and 2. Each trit can be encoded in two bits, but two bits can have four values. Thus, it makes sense to work with a ternary number system where each trit is represented by two bits and in addition to the three trits there is a fourth symbol, a comma (c). Once we include the c, it becomes easy to construct the ternary comma code for the integers. The comma code of n is simply the ternary representation of n - 1 followed by a c. Thus, the comma code of 8 is 21c (because 7 = 2 · 3+1) and the comma code of 18 is 122c (because 17 = 1 ·9+2· 3+2). Table 3.22 (after [Fenwick 96a]) lists several ternary comma codes (the columns labeled L are the length of the code, in bits). These codes start long (longer than most of the other codes described here) but grow slowly. Thus, they are suitable for applications where large integers are common. These codes are also easy to encode and decode and their principal downside is the comma symbol (signalling the end of a code) that requires two bits. This inefficiency is not serious, but becomes more so for comma codes based on larger number bases. In a base-15 comma code, for example, each of the 15 digits requires four bits and the comma is also a 4-bit pattern. Each code ends with a 4-bit comma, instead of with the theoretical minimum of one bit, and this feature renders such codes inefficient. (However, the overall redundancy per symbol decreases for large number bases. In a base-7 system, one of eight symbols is sacrificed for the comma, while in a base 15 it is one of 16 symbols.) The ideas above are simple and easy to implement, but they make sense only for long bitstrings. In data compression, as well as in many other applications, we normally need short codes, ranging in size from two to perhaps 12 bits. Because they are used for compression, such codes have to be self-delimiting without adding any extra bits. The 3.11 Location Based Encoding (LBE) 117 Value Code L Value Code L 0 c 2 11 101c 8 1 0c 4 12 102c 8 2 1c 4 13 110c 8 3 2c 4 14 111c 8 4 10c 6 15 112c 8 5 11c 6 16 120c 8 6 12c 6 17 121c 8 7 20c 6 18 122c 8 8 21c 6 19 200c 8 9 22c 6 20 201c 8 . . . . . . 64 2100c 10 1,000 1101000c 16 128 11201c 12 3,000 11010002c 18 256 100110c 14 10,000 111201100c 20 512 200221c 14 65,536 10022220020c 24 Table 3.22: Ternary Comma Codes and Their Lengths. solution is to design sets of prefix codes, codes that have the prefix property. Among these codes, the ones that yield the best performance are the universal codes. In 1840, Thomas Fowler, a self-taught English mathematician and inventor, created a unique ternary calculating machine. Until recently, all detail of this machine was lost. A research project begun in 1997 uncovered sufficient information to enable the recreation of a physical concept model of Fowler’s machine. The next step is to create a historically accurate replica. —[Glusker et al 05] 3.11 Location Based Encoding (LBE) Location based encoding (LBE) is a simple method for assigning variable-length codes to a set of symbols, not necessarily integers. The originators of this method are P. S. Chitaranjan, Arun Shankar, and K. Niyant [LBE 07]. The LBE encoder is essentially a two-pass algorithm. Given a data file to be compressed, the encoder starts by reading the file and counting symbol frequencies. Following the first pass, it (1) sorts the symbols in descending order of their probabilities, (2) stores them in a special order in a matrix, and (3) assigns them variable-length codes. In between the passes, the encoder writes the symbols, in their order in the matrix, on the compressed stream, for the use of the decoder. The second pass simply reads the input file again symbol by symbol and replaces each symbol with its code. The decoder starts by reading the sorted sequence of symbols from the compressed stream and constructing the codes in lockstep with the encoder. The decoder then reads the codes off its input and replaces each code with the original symbol. Thus, decoding is fast. 118 3. Advanced VL Codes The main idea is to assign short codes to the common symbols by placing the symbols in a matrix in a special diagonal order, vaguely reminiscent of the zigzag order used by JPEG, such that high-probability symbols are concentrated at the top-left corner of the matrix. This is illustrated in Figure 3.23a, where the numbers 1, 2, 3, . . . indicate the most-common symbol, the second most-common one, and so on. Each symbol is assigned a variable-length code that indicates its position (row and column) in the matrix. Thus, the code of 1 (the most-common symbol) is 11 (row 1 column 1), the code of 6 is 0011 (row 3 column 1), and the code of 7 is 10001 (row 1 column 4). Each code has two 1’s, which makes it trivial for the decoder to read and identify the codes. The length of a code depends on the diagonal (shown in gray in the figure), and the figure shows codes with lengths (also indicated in gray) from two to eight bits. 1 2 2 4 4 7 7 11 16 3 3 5 5 8 8 11 1111 12 17 22 6 6 9 13 18 10 14 19 15 20 21 28 - 1 3 6 12 20 2 4 7 13 21 33 32 5 8 14 22 34 9 15 23 35 16 24 36 25 37 38 - 10 17 26 39 11 18 27 40 19 28 41 29 42 43 - 30 44 31 45 46 (a) (b) T1 T2 T3 Figure 3.23: LBE Code Tables. We know that Huffman coding fails for very small alphabets. For a 2-symbol alphabet, the Huffman algorithm assigns 1-bit codes to the two symbols regardless of their probabilities, so no compression is achieved. Given the same alphabet, LBE assigns the two codes 11 and 101 to the symbols, so the average length of its codes is between 2 and 3 (the shorter code is assigned to the most-common symbol, which is why the average code length is in the interval (2, 2.5], still bad). Thus, LBE performs even worse in this case. For alphabets with small numbers of symbols, such as three or four, LBE performs better, but still not as good as Huffman. The ASCII code assigns 7-bit codes to 128 symbols, and this should be compared with the average code length achievable by LBE. LBE assigns codes of seven or fewer bits to the 21 most-common symbols. Thus, if the alphabet consists of 21 or fewer symbols, replacing them with LBE codes generates compression. For larger alphabets, longer codes are needed, which may cause expansion (i.e., compression where the average code length is greater than seven bits). It is possible to improve LBE’s performance somewhat by constructing several tables, as illustrated in Figure 3.23b. Each code is preceded by its table number such that the codes of table n are preceded by (2n-2) 1’s. The codes of table 1 are not preceded by any 1’s, those in table 2 are preceded by 11, those of table 3 are preceded by 1111, and so on. Because of the presence of pairs of 1’s, position (1, 1) of a table cannot be used and the first symbol of a table is placed at position (1, 2). The codes are placed in table 1 until a code can become shorter if placed in table 2. This happens with the 10th 3.12 Stout Codes 119 most-common symbol. Placing it in table 1 would have assigned it the code 100001, whereas placing it in table 2 (which is empty so far) assigns it the shorter code 11|101. The next example is the 12th symbol. Placing it in table 2 would have assigned it code 11|1001, but placing it in table 1 gives it the same-length code 100001. With this scheme it is possible to assign short codes (seven bits or fewer) to the 31 most-common symbols. Notice that there is no need to actually construct any tables. Both encoder and decoder can determine the codes one by one simply by keeping track of the next available position in each table. Once a code has been computed, it can be stored with other codes in an array or another convenient data structure. The decoder reads pairs of consecutive 1’s until it reaches a 0 or a single 1. This provides the table number, and the rest of the code is read until two 1’s (not necessarily consecutive) are found. The method is simple but not very efficient, especially because the compressed stream has to start with a list of the symbols in ascending order of probabilities. The main advantage of LBE is simple implementation, simple encoding, and fast decoding. Die Mathematiker sind eine Art Franzosen: redet man zu ihnen, so uebersetzen sie es in ihre Sprache, und dann ist es also bald ganz etwas anderes. (The mathematicians are a sort of Frenchmen: when you talk to them, they immediately translate it into their own language, and right away it is something entirely different.) —Johann Wolfgang von Goethe 3.12 Stout Codes In his 1980 short paper [Stout 80], Quentin Stout introduced two families Rl and Sl of recursive, variable-length codes for the integers, similar to and more general than Elias omega and Even–Rodeh codes. The Stout codes are universal and asymptotically optimal. Before reading ahead, the reader is reminded that the length L of the integer n is given by L = 1+log2 n (Equation (2.1)). The two families of codes depend on the choice of an integer parameter l. Once a value (greater than or equal to 2) for l has been selected, a codeword in the first family consists of a prefix Rl(n) and a suffix 0n. The suffix is the binary value of n in L bits, preceded by a single 0 separator. The prefix Rl(n) consists of length groups. It starts with the length L of n, to which is prepended the length L1 of L, then the length L2 of L1, and so on until a length Li is reached that is short enough to fit in an l-bit group. Notice that the length groups (except perhaps the leftmost one) start with 1. Thus, the single 0 separator indicates the end of the length groups. As an example, we select l = 2 and encode a 985-bit integer n whose precise value is irrelevant. The suffix is the 986-bit string 0n and the prefix starts with the group L = 985 = 11110110012. The length L1 of this group is 1010 = 10102, the length L2 of the second group is 4 = 1002, and the final length L3 is 3 = 112. The complete codeword is therefore 11|100|1010|1111011001|0|n. Notice how the length groups start 120 3. Advanced VL Codes with 1, which implies that each group is followed by a 1, except the last length group which is followed by a 0. Decoding is simple. The decoder starts by reading the first l bits. These are the length of the next group. More and more length groups are read, until a group is found that is followed by a 0. This indicates that the next group is n itself. With this background, the recursive definition of the Rl prefixes is easy to read and understand. We use the notation L = 1+log2 n and denote by B(n, l) the l-bit binary representation (beta code) of the integer n. Thus, B(12, 5) = 01100. For l . 2, the prefixes are defined by Rl(n) = B(n, l), for 0 . n . 2l - 1, Rl(L)B(n,L), for n . 2l. Those familiar with the Even–Rodeh code (Section 3.7) may already have realized that this code is identical to R3. Furthermore, the Elias omega code (Section 3.4) is in-between R2 and R3 with two differences: (1) the omega code encodes the quantities Li - 1 and (2) the 0 separator is placed to the right of n. R2(985) = 11 100 1010 1111011001 R3(985) = 100 1010 1111011001 R4(985) = 1010 1111011001 R5(985) = 01010 1111011001 R6(985) = 001010 1111011001 R2(31,925) = 11 100 1111 111110010110101 R3(31,925) = 100 1111 111110010110101 R4(31,925) = 1111 111110010110101 R5(31,925) = 01111 111110010110101 R6(31,925) = 001111 111110010110101 The short table lists the Rl(n) prefixes for 2 . l . 6 and 985-bit and 31,925-bit integers. It is obvious that the shortest prefixes are obtained for parameter values l = L = 1+log2 n. Larger values of l result in slightly longer prefixes, while smaller values require more length groups and therefore result in much longer prefixes. Elementary algebraic manipulations indicate that the range of best lengths L for a given parameter l is given by [2s, 2e - 1] where s = 2l-1 and e = 2l - 1 = 2s - 1 (this is the range of best lengths L, not best integers n). The following table lists these intervals for a few l values and makes it clear that small parameters, such as 2 and 3, are sufficient for most practical applications of data compression. Parameter l = 2 is best for integers that are 2 to 7 bits long, while l = 3 is best for integers that are 8 to 127 bits long. l s e 2s 2e - 1 2 2 3 2 7 3 4 7 8 127 4 8 15 128 32,767 5 9 31 32,768 2,147,483,647 The second family of Stout codes is similar, but with different prefixes that are denoted by Sl(n). For small l values, this family offers some improvement over the Rl codes. Specifically, it removes the slight redundancy that exists in the Rl codes because a length group cannot be 0 (which is why a length group in the omega code encodes Li-1 and not Li). The Sl prefixes are similar to the Rl prefixes with the difference that a length group for Li encodes Li - 1 - l. Thus, S2(985), the prefix of a 985-bit integer n, starts with the 10-bit length group 1111011001 and prepends to it the length group 3.12 Stout Codes 121 for 10-1-2 = 7 = 1112. To this is prepended the length group for 3-1-2 = 0 as the two bits 00. The result is the 15-bit prefix 00|111|1111011001, shorter than the 19 bits of R2(985). Another example is S3(985), which starts with the same 1111011001 and prepends to it the length group for 10 - 1 - 3 = 6 = 1102. The recursion stops at this point because 110 is an l-bit group. The result is the 13-bit codeword 110|1111011001, again shorter than the 17 bits of R3(985). The Sl(n) prefixes are defined recursively by Sl(n) = B(n, l), for 0 . n . 2l - 1, Rl(L - 1 - l)B(n,L), for n . 2l. Table 3.24 lists some S2(n) and S3(n) prefixes and illustrates their regularity. Notice that the leftmost column lists the values of L, i.e., the lengths of the integers being encoded, and not the integers themselves. A length group retains its value until the group that follows it becomes all 1’s, at which point the group increments by 1 and the group that follows is reset to 10 . . . 0. All the length groups, except perhaps the leftmost one, start with 1. This regular behavior is the result of the choice Li - 1 - l. L S2(n) S3(n) 1 01 001 2 10 010 3 11 011 4 00 100 100 5 00 101 101 6 00 110 110 7 00 111 111 8 01 1000 000 1000 15 01 1111 000 1111 16 10 10000 001 10000 32 11 100000 010 100000 64 00 100 1000000 011 1000000 128 00 101 10000000 100 10000000 256 00 110 100000000 101 100000000 512 00 111 1000000000 110 1000000000 1024 01 1000 10000000000 111 10000000000 2048 01 1001 100000000000 000 1000 100000000000 Table 3.24: S2(n) and S3(n) Codes. The prefix S2(64), for example, starts with the 7-bit group 1000000 = 64 and prepends to it S2(7 - 1 - 2) = S2(4) = 00|100. We stress again that the table lists only the prefixes, not the complete codewords. Once this is understood, it is not hard to see that the second Stout code is a prefix code. Once a codeword has been assigned, it will not be the prefix of any other codeword. Thus, for example, the prefixes of all the codewords for 64-bit integers start with the prefix 00 100 of the 4-bit integers, but any codeword for a 4-bit integer has a 0 following the 00 100, whereas the codewords for the 64-bit integers have a 1 following 00 100. 122 3. Advanced VL Codes 3.13 Boldi–Vigna (.) Codes The World Wide Web (WWW) was created in 1990 and currently, after only 20 years of existence, it completely pervades our lives. It is used by many people and is steadily growing and finding more applications. Formally, the WWW is a collection of interlinked, hypertext documents (web pages) that can be sent over the Internet. Mathematically, the pages and hyperlinks may be viewed as nodes and edges in a vast directed graph (a webgraph) which itself has become an important object of study. Currently, the graph of the entireWWWconsists of hundreds of millions of nodes and over a billion links. Experts often claim that these numbers grow exponentially with time. The various web search engines crawl over the WWW, collecting pages, resolving their hypertext links, and creating large webgraphs. Thus, a search engine operates on a vast data base and it is natural to want to compress this data. One approach to compressing a webgraph is outlined in [Randall et al. 01]. Each URL is replaced by an integer. These integers are stored in a list that includes, for each integer, a pointer to an adjacency list of links. Thus, for example, URL 104 may be a web page that links to URLs 101, 132, and 174. These three integers become the adjacency list of 104, and this list is compressed by computing the differences of adjacent links and replacing the differences with appropriate variable-length codes. In our example, the differences are 101 - 104 = -3, 132 - 101 = 31, and 174 - 132 = 42. Experiments with large webgraphs indicate that the differences (also called gaps or deltas) computed for a large webgraph tend to be distributed according to a power law whose parameter in normally in the interval [1.1, 1.3], so the problem of compressing a webgraph is reduced to the problem of finding a variable-length code that corresponds to such a distribution. A power law distribution has the form Z.[n] = P n. where Z.[n] is the distribution (the number of occurrences of the value n), . is a parameter of the distribution, and P is a constant of proportionality. Codes for power law distributions have already been mentioned. The length of the Elias gamma code (Equation (3.1)) is 2log2 n+1, which makes it a natural choice for a distribution of the form 1/(2n2). This is a power law distribution with . = 2 and P = 1/2. Similarly, the Elias delta code is suitable for a distribution of the form 1/[2n(log2(2n))2]. This is very close to a power law distribution with . = 1. The nibble code and its variations (page 92) correspond to power law distributions of the form 1/n1+ 1 k , where the parameter is 1+ 1 k . The remainder of this section describes the zeta (.) code, also known as Boldi–Vigna code, introduced by Paolo Boldi and Sebastiano Vigna as a family of variable-length codes that are best choices for the compression of webgraphs. The original references are [Boldi and Vigna 04a] and [Boldi and Vigna 04b]. The latest reference is [Boldi and Vigna 05]. We start with Zipf’s law, an empirical power law [Zipf 07] introduced by the linguist George K. Zipf. It states that the frequency of any word in a natural language is roughly inversely proportional to its position in the frequency table. Intuitively, Zipf’s law states 3.13 Boldi–Vigna (.) Codes 123 that the most frequent word in any language is twice as frequent as the second mostfrequent word, which in turn is twice as frequent as the third word, and so on. For a language having N words, the Zipf power law distribution with a parameter . is given by Z.[n] = 1/n. N i=1 1 i. If the set is infinite, the denominator becomes the well-known Riemann zeta function .(.) = . i=1 1 i. which converges to a finite value for . > 1. In this case, the distribution can be written in the form Z.[n] = 1 .(.)n. The Boldi–Vigna zeta code starts with a positive integer k that becomes the shrinking factor of the code. The set of all positive integers is partitioned into the intervals [20, 2k - 1], [2k, 22k - 1], [22k, 23k - 1], and in general [2hk, 2(h+1)k - 1]. The length of each interval is 2(h+1)k - 2hk. Next, a minimal binary code is defined, which is closely related to the phased-in codes of Section 2.9. Given an interval [0, z-1] and an integer x in this interval, we first compute s = log2 z. If x < 2s - z, it is coded as the xth element of the interval, in s - 1 bits. Otherwise, it is coded as the (x - z - 2s)th element of the interval in s bits. With this background, here is how the zeta code is constructed. Given a positive integer n to be encoded, we employ k to determine the interval where n is located. Once this is known, the values of h and k are used in a simple way to construct the zeta code of n in two parts, the value of h + 1 in unary (as h zeros followed by a 1), followed by the minimal binary code of n - 2hk in the interval [0, 2(h+1)k - 2hk - 1]. Example. Given k = 3 and n = 16, we first determine that n is located in the interval [23, 26 -1], which corresponds to h = 1. Thus, h+1 = 2 and the unary code of 2 is 01. The minimal binary code of 16 - 23 = 8 is constructed in steps as follows. The length z of the interval [23, 26 -1] is 56. This implies that s = log2 56 = 6. The value 8 to be encoded satisfies 8 = 26 - 56, so it is encoded as x - z - 2s = 8- 56 - 26 = 16 in six bits, resulting in 010000. Thus, the .3 code of n = 16 is 01|010000. Table 3.25 lists . codes for various shrinking factors k. For k = 1, the .1 code is identical to the . code. The nibble code of page 92 is also shown. A nibble (more accurately, nybble) is the popular term for a 4-bit number (or half a byte). A nibble can therefore have 16 possible values, which makes it identical to a single hexadecimal digit The length of the zeta code of n with shrinking factor k is 1+(log2 n)/k(k+1)+ . (n) where . (n) = 0, if (log2 n)/k - (log2 n)/k . [0, 1/k), 1, otherwise. 124 3. Advanced VL Codes n .= .1 .2 .3 .4 . Nibble 1 1 10 100 1000 1 1000 2 010 110 1010 10010 0100 1001 3 011 111 1011 10011 0101 1010 4 00100 01000 1100 10100 01100 1011 5 00101 01001 1101 10101 01101 1100 6 00110 01010 1110 10110 01110 1101 7 00111 01011 1111 10111 01111 1110 8 0001000 011000 0100000 11000 00100000 1111 9 0001001 011001 0100001 11001 00100001 00011000 10 0001010 011010 0100010 11010 00100010 00011001 11 0001011 011011 0100011 11011 00100011 00011010 12 0001100 011100 0100100 11100 00100100 00011011 13 0001101 011101 0100101 11101 00100101 00011100 14 0001110 011110 0100110 11110 00100110 00011101 15 0001111 011111 0100111 11111 00100111 00011110 16 000010000 00100000 01010000 010000111 001011001 00011111 Table 3.25: .-Codes for 1 . k . 4 Compared to ., . and Nibble Codes. Thus, this code is ideal for integers n that are distributed as 1 + . (n) n1+ 1 k . This is very close to a power law distribution with a parameter 1 + 1 k . The developers show that the zeta code is complete, and they provide a detailed analysis of the expected lengths of the zeta code for various values of k. The final results are summarized in Table 3.26 where certain codes are recommended for various ranges of the parameter . of the distribution. . : < 1.06 [1.06, 1.08] [1.08, 1.11] [1.11, 1.16] [1.16, 1.27] [1.27, 1.57] [1.57, 2] Code: . .6 .5 .4 .3 .2 . = .1 Table 3.26: Recommended Ranges for Codes. 3.14 Yamamoto’s Recursive Code 125 3.14 Yamamoto’s Recursive Code In 2000, Hirosuke Yamamoto came up with a simple, ingenious way [Yamamoto 00] to improve the Elias omega code. As a short refresher on the omega code, we start with the relevant paragraph from Section 3.4. The omega code uses itself recursively to encode the prefix M, which is why it is sometimes referred to as a recursive Elias code. The main idea is to prepend the length of n to n as a group of bits that starts with a 1, then prepend the length of the length, as another group, to the result, and continue prepending lengths until the last length is 2 or 3 (and it fits in two bits). In order to distinguish between a length group and the last, rightmost group (that of n itself), the latter is followed by a delimiter of 0, while each length group starts with a 1. The decoder of the omega code reads the first two bits (the leftmost length group), interprets its value as the length of the next group, and continues reading groups until a group is found that is followed by the 0 delimiter. That group is the binary representation of n. This scheme makes it easy to decode the omega code, but is somewhat wasteful, because the MSB of each length group is 1, and so the value of this bit is known in advance and it acts only as a separator. Yamamoto’s idea was to select an f-bit delimiter a (where f is a small positive integer, typically 2 or 3) and use it as the delimiter instead of a single 0. In order to obtain a UD code, none of the length groups should start with a. Thus, the length groups should be encoded with special variable-length codes that do not start with a. Once a has been selected and its length f is known, the first step is to prepare all the binary strings that do not start with a and assign each string as the code ˜B a,f () of one of the positive integers. Now, if a length group has the value n, then ˜B a,f (n) is placed in the codeword instead of the binary value of n. We start with a simpler code that is denoted by Ba,f (n). As an example, we select f = 2 and the 2-bit delimiter a = 00. The binary strings that do not start with 00 are the following: the two 1-bit strings 0 and 1; the three 2-bit strings 01, 10, and 11; the six 3-bit strings 010, 011, 100, 101, 110, and 111; the 12 4-bit strings 0100, 1001, through 1111 (i.e., the 16 4-bit strings minus the four that start with 00), and so on. These strings are assigned to the positive integers as listed in the second column of Table 3.27. The third column of this table lists the first 26 B codes for a = 100. One more step is needed. Even though none of the Ba,f (n) codes starts with a, it may happen that a short Ba,f (n) code (a code whose length is less than f) followed by another Ba,f (n) code will accidentally contain the pattern a. Thus, for example, the table shows that B100,3(5) = 10 followed by B100,3(7) = 000 becomes the string 10|000 and may be interpreted by the decoder as 100|00 . . .. Thus, the B100,3(n) codes have to be made UD by eliminating some of the short codes, those that are shorter than f bits and have the values 1 or 10. Similarly, the single short B00,2(1) = 0 code should be discarded. The resulting codes are designated ˜B a,f (n) and are listed in the last two columns of the table. If the integers to be encoded are known not to exceed a few hundred or a few thousand, then both encoder and decoder can have built-in tables of the ˜B a,f (n) codes 126 3. Advanced VL Codes n Ba,f (n) ˜B a,f (n) a = 00 a = 100 a = 00 a = 100 1 0 0 1 0 2 1 1 01 00 3 01 00 10 01 4 10 01 11 11 5 11 10 010 000 6 010 11 011 001 7 011 000 100 010 8 100 001 101 011 9 101 010 110 101 10 110 011 111 110 11 111 101 0100 111 12 0100 110 0101 0000 13 0101 111 0110 0001 14 0110 0000 0111 0010 15 0111 0001 1000 0011 16 1000 0010 1001 0100 17 1001 0011 1010 0101 18 1010 0100 1011 0110 19 1011 0101 1100 0111 20 1100 0110 1101 1010 21 1101 0111 1110 1011 22 1110 1010 1111 1100 23 1111 1011 01000 1101 24 01000 1100 01001 1110 25 01001 1101 01010 1111 26 01010 1110 01011 00000 Table 3.27: Some Ba,f (n) and ˜B a,f (n) Yamamoto Codes. for all the relevant integers. In general, these codes have to be computed “on the fly” and the developer provides the following expression for them (the notation K[j] means the integer K expressed in j bits, where some of the leftmost bits may be zeros). ˜B a,f (n) =.. ... .. [n -M(j, f) + L(j, f)][j], if M(j, f) - L(j, f) . n < M(j, f) - L(j, f) + N(j, f, a), [n -M(j, f - 1) + L(j + 1, f)][j], if M(j, f) - L(j, f) + N(j, f, a) . n < M(j + 1, f) - L(j + 1, f), where M(j, f) = 2j - 2(j-f)+, (t)+ = max(t, 0), L(j, f) = (f - 1) - (f - j)+, and N(j, f, a) = 2j-fa. Given a positive integer n, the first step in encoding it is to determine the value of j by examining the inequalities. Once j is known, functions M and L can be computed. Armed with the ˜B a,f (n) codes, the recursive Yamamoto code Ca,f (n) is easy to describe. We select an f-bit delimiter a. Given a positive integer n, we start with the 3.14 Yamamoto’s Recursive Code 127 group ˜B a,f (n). Assuming that this group is n1 bits long, we prepend to it the length group ˜B a,f (n1 - 1). If this group is n2 bits long, we prepend to it the length group ˜B a,f (n2 - 1). This is repeated recursively until the last length group is ˜B a,f (1). (This code is always a single bit because it depends only on a, and the decoder knows this bit because it knows a. Thus, in principle, the last length group can be omitted, thereby shortening the codeword by one bit.) The codeword is completed by appending the delimiter a to the right end. Table 3.28 lists some codewords for a = 00 and for a = 100. n a= 00 a = 100 1 1 00 0 100 2 1 01 00 0 00 100 3 1 10 00 0 01 100 4 1 11 00 0 11 100 5 1 01 010 00 0 00 000 100 6 1 01 011 00 0 00 001 100 7 1 01 100 00 0 00 010 100 8 1 01 101 00 0 00 011 100 9 1 01 110 00 0 00 101 100 10 1 01 111 00 0 00 110 100 11 1 10 0100 00 0 00 111 100 12 1 10 0101 00 0 01 0000 100 13 1 10 0110 00 0 01 0001 100 14 1 10 0111 00 0 01 0010 100 15 1 10 1000 00 0 01 0011 100 16 1 10 1001 00 0 01 0100 100 17 1 10 1010 00 0 01 0101 100 18 1 10 1011 00 0 01 0110 100 19 1 10 1100 00 0 01 0111 100 20 1 10 1101 00 0 01 1010 100 21 1 10 1110 00 0 01 1011 100 22 1 10 1111 00 0 01 1100 100 23 1 11 01000 00 0 01 1101 100 24 1 11 01001 00 0 01 1110 100 25 1 11 01010 00 0 01 1111 100 26 1 11 01011 00 0 11 00000 100 Table 3.28: Yamamoto Recursive Codewords. The developer, Hirosuke Yamamoto, also proves that the length of a codeword Ca,f (n) is always less than or equal to log*2 n + F(f)wf (n) + cf + .(n). More importantly, for infinitely many integers, the length is also less than or equal to log*2 n + (1 - F(f))wf (n) + cf + 2.(n), where F(f) = -log2(1 - 2-f ), cf = 5(f - 2)+ + f + 5F(f), .(n) . log2 e n 1 + (w(n) - 1)(log2 e)w(n)-1 log2 n . 4.7 n , 128 3. Advanced VL Codes and the log-star function log*2 n is the finite sum log2 n + log2 log2 n + · · · + log2 log2 . . . log2    w(n) n where w(n) is the largest integer for which the compound logarithm is still nonnegative. The important point (at least from a theoretical point of view) is that this recursive code is shorter than log*2 n for infinitely many integers. This code can also be extended to other number bases and is not limited to base-2 numbers. 3.15 VLCs and Search Trees There is an interesting association between certain unbounded searches and prefix codes. A search is a classic problem in computer science. Given a data structure, such as an array, a list, a tree, or a graph, where each node contains an item, the problem is to search the structure and find a given item in the smallest possible number of steps. The following are practical examples of computer searches. 1. A two-player game. Player A chooses a positive integer n and player B has to guess n by asking only the following type of question: Is the integer i less than n? Clearly, the number of questions needed to find n depends on n and on the strategy (algorithm) used by B. 2. Given a function y = f(x), find all the values of x for which the function equals a given value Y (typically zero). 3. An online business stocks many items. Customers should be able to search for an item by its name, price range, or manufacturer. The items are the nodes of a large structure (normally a tree), and there must be a fast search algorithm that can find any item among many thousands of items in a few seconds. 4. An Internet search engine faces a similar problem. The first part of such an engine is a crawler that locates a vast number of Web sites, inputs them into a data base, and indexes every term found. The second part is a search algorithm that should be able to locate any term (out of many millions of terms) in just a few seconds. (There is also the problem of ranking the results.) Internet users search for terms (such as “variable-length codes”) and naturally prefer search engines that provide results in just a few seconds. Searching a data base, even a very large one, is considered a bounded search, because the search space is finite. Bounded searches have been a popular field of research for many years, and many efficient search methods are known and have been analyzed in detail. In contrast, searching for a zero of a function is an example of unbounded search, because the domain of the function is normally infinite. In their work [Bentley and Yao 76], Jon Bentley and Andrew Yao propose several algorithms for searching a linearly ordered unbounded table and show that each algorithm is associated with a binary string that can be considered a codeword in a prefix code. This creates an interesting and unexpected relation between searching and variable-length codes, two hitherto unrelated areas of scientific endeavor. The authors restrict themselves to algorithms that search for an integer n in an ordered table F(i) using only elementary tests of the form F(i) < n. 3.15 VLCs and Search Trees 129 The simplest search method, for both bounded and unbounded searches, is linear search. Given an ordered array of integers F(i) and an integer n, a linear search performs the tests F(1) < n, F(2) < n,. . . until a match is found. The answers to the tests are No, No,. . . , Yes, which can be expressed as the bitstring 11 . . . 10. Thus, a linear search (referred to by its developers as algorithm B0) corresponds to the unary code. The next search algorithm (designated B1) is an unbounded variation of the wellknown binary search. The first step of this version is a linear search that determines an interval. This step starts with an interval [F(a), F(b)] and performs the test F(b) < n. If the answer is No, then n is in this interval, and the algorithm executes its second step, where n is located in this interval with a bounded binary search. If the answer is Yes, then the algorithm computes another interval that starts at F(b + 1) and is twice as wide as the previous interval. Specifically, the algorithm computes F(2i - 1) for i = 1, 2,. . . and each interval is [F(2i-1), F(2i - 1)]. Figure 3.29 (compare with Figures 2.15 and 2.16) shows how these intervals can be represented as a binary search tree where the root and each of its right descendants correspond to an index 2i - 1. The left son of each descendant is the root of a subtree that contains the remaining indexes less than or equal to 2i-1. Thus, the left son of the root is index 1. The left son of node 3 is the root of a subtree containing the remaining indexes that are less than or equal to 3, namely 2 and 3. The left son of 7 is the root of a subtree with 4, 5, 6, and 7, and so on. 1 1 1 1 0 0 0 1 1 2 2 3 3 4 4 5 5 6 6 7 7 9 13 8 10 12 14 8 9 10 11 11 12 13 14 15 15 Figure 3.29: Gamma Code and Binary Search. The figure also illustrates how n = 12 is searched for and located by the B1 method. The algorithm compares 12 to 1, 3, 7, and 15 (a linear search), with results No, No, No, and Yes, or 1110. The next step performs a binary search to locate 12 in the interval [8, 15] with results 100. These results are listed as labels on the appropriate edges. The final bitstring associated with 12 by this method is therefore 1110|100. This result is the Elias gamma code of 12, as produced by the alternative construction method of page 101 (this construction produces codes different from those of Table 3.10, but the code lengths are the same). If m is the largest integer such that 2m-1 . n < 2m, then the first step (a linear search to determine m) requires m = log2 n + 1 tests. The second step requires log2 2m-1 = m-1 = 2log2 n tests. The total number of tests is therefore 2log2 n+1, which is the length of the gamma code (Equation (3.1)). The third search method proposed by Bentley and Yao (their algorithm B2) is a double binary search. The second step is the same, but the first step is modified to determine m by constructing intervals that become longer and longer. The intervals 130 3. Advanced VL Codes computed by method B1 start at 2i - 1 for i = 1, 2, 3, . . . , or 1, 3, 7, 15, 31, . . . . Method B2 constructs intervals that start with numbers of the form 22i-1 -1 for i = 1, 2, 3, . . . . These numbers are 1, 7, 127, 32,767, 231 - 1, and so on. They become the roots of subtrees as shown in Figure 3.30. There are fewer search trees, which speeds up step 1, but they grow very fast. The developers show that the first step requires 1 + 2log2 n tests (compared to the m = log2 n + 1 tests required by B1) and the second step requires log2 n tests (same as B1), for a total of log2 n + 2log2(log2 n + 1) + 1 = 1+log2 n + 2log2 log2(2n). The number of tests is identical to the length of the Elias delta code, Equation (3.2). Thus, the variable-length code associated with this search tree is a variant of the delta code. The codewords are different from those listed in Table 3.12, but the code is equivalent because of the identical lengths. 1 1 4 4 5 5 2 2 3 4 5 7 16 64 12 24 48 86 8 9 10 11 32 32 63 64 127 127 32767 6 6 7 6 9 8 10 12 13 14 15 13 12 14 12 24 25 26 27 25 24 26 28 29 30 31 29 28 30 27 16 17 18 19 17 16 18 20 21 22 23 21 20 22 19 2 31-1 Figure 3.30: Delta Code and Binary Search. Bentley and Yao go on to propose unbounded search algorithms Bk (k-nested binary search) and U (ultimate). The latter is shown by [Ahlswede et al. 97] to be associated with a modified Elias omega code. The authors of this paper also construct the binary search trees for the Stout codes (Section 3.12). 3.16 Taboo Codes 131 3.16 Taboo Codes The taboo approach to variable-length codes, as well as the use of the term “taboo,” are the brainchilds of Steven Pigeon. The two types of taboo codes are described in [Pigeon 01a,b] where it is shown that they are universal (more accurately, they can be made as close to universal as desired by the choice of a parameter). The principle of the taboo codes is to select a positive integer parameter n and reserve a pattern of n bits to indicate the end of a code. This pattern should not appear in the code itself, which is the reason for the term taboo. Thus, the taboo codes can be considered suffix codes. The first type of taboo code is block-based and its length is a multiple of n. The block-based taboo code of an integer is a string of n-bit blocks, where n is a user-selected parameter and the last block is a taboo bit pattern that cannot appear in any of the other blocks. An n-bit block can have 2n values, so if one value is reserved for the taboo pattern, each of the remaining code blocks can have one of the remaining 2n - 1 bit patterns. In the second type of taboo codes, the total length of the code is not restricted to a multiple of n. This type is called unconstrained and is shown to be related to the n-step Fibonacci numbers. We use the notation n: t to denote a string of n bits concatenated with the taboo string t. Table 3.31 lists the lengths, number of codes, and code ranges as the codes get longer when more blocks are added. Each row in this table summarizes the properties of a range of codes. The number of codes in the kth range is (2n-1)k, and the total number of codes gn(k) in the first k ranges is obtained as the sum of a geometric progression gn(k) = k i=1 (2n - 1)i = [(2n - 1)k - 1](2n - 1) 2n - 2 . Codes Length # of values Range n: t 2n 2n - 1 0 to (2n - 1) - 1 2n: t 3n (2n - 1)2 (2n - 1) to (2n - 1) + (2n - 1)2 - 1 ... ... kn: t (k + 1)n (2n - 1)k k-1 i=1 (2n - 1)i to -1 +k i=1(2n - 1)i ... ... Table 3.31: Structure of Block-Based Taboo Codes. The case n = 1 is special and uninteresting. A 1-bit block can either be a 0 or a 1. If we reserve 0 as the taboo pattern, then all the other blocks of codes cannot contain any zeros and must be all 1’s. The result is the infinite set of codewords 10, 110, 1110,. . . , which is the unary code and therefore not universal (more accurately, the unary code is .-universal). The first interesting case is n = 2. Table 3.32 lists some codes for this 132 3. Advanced VL Codes m Code m Code m Code m Code 0 0100 4 011000 8 101100 12 01 01 01 00 1 1000 5 011100 9 110100 13 01 01 10 00 2 1100 6 100100 10 11 10 00 14 01 01 11 00 3 010100 7 101000 11 11 11 00 . . . Table 3.32: Some Block-Based Taboo Codes for n = 2. case. Notice that the pattern of all zeros is reserved for the taboo, but any other n-bit pattern would do as well. In order to use this code in practice, we need simple, fast algorithms to encode and decode it. Given an integer m, we first determine the number of n-bit blocks in the code of m. Assuming that m appears in the kth range of Table 3.31, we can write the basic relation m . gn(k) - 1. This can be written explicitly as m . [(2n - 1)k - 1](2n - 1) 2n - 2 - 1 or k . log2 m+ 2 - m+1 2n-1  log2(2n - 1) . The requirement that m be less than gn(k) yields the value of k (which depends on both n and m) as kn(m) =.... log2 m+ 2 - m+1 2n-1  log2(2n - 1) ..... Thus, the code of m consists of kn(m) blocks of n bits each plus the n-bit taboo block, for a total of [kn(m) + 1]n bits. This code can be considered the value of the integer c def = m- gn(kn(m) - 1) represented in base (2n - 1), where each “digit” is n bits long. Encoding: Armed with the values of n, m, and c, we can write the individual values of the n-bit blocks bi that constitute the code of m as bi = c(2n - 1)i mod (2n - 1)+ 1, i= 0, 1, . . . , kn(m) - 1. (3.3) The taboo pattern is the (kn(m) + 1)th block, block bkn(m). Equation (3.3) is the basic relation for encoding integer m. Decoding: The decoder knows the value of n and the taboo pattern. It reads n-bit blocks bi until it finds the taboo block. The number of blocks read is k (rather kn(m)) and once k is known, decoding is done by m = gn(k - 1) + k-1 i=0 (bi - 1)(2n - 1)i. (3.4) 3.16 Taboo Codes 133 Equations (3.3) and (3.4) lend themselves to fast, efficient implementation. In particular, decoding is easy. The decoder can read n-bit blocks bi and add terms to the partial sum of Equation (3.4) until it reaches the taboo block, at which point it only has to compute and add gn kn(m) - 1 to the sum in order to obtain m. The block-based taboo code is longer than the binary representation (beta code) of the integers by n+kn-log2(2n -1) bits. For all values of n except the smallest ones, 2n -1 . 2n, so this difference of lengths (the “waste” of the code) equals approximately n. Even for n = 4, the waste is only n + 0.093k. Thus, for large integers m, the waste is very small compared to m. The developer of these codes, Steven Pigeon, proves that the block-based taboo code can be made as close to universal as desired. Let us overthrow the totems, break the taboos. Or better, let us consider them cancelled. Coldly, let us be intelligent. —Pierre Trudeau Now, for the unconstrained taboo codes. The idea is to have a bitstring of length l, the actual code, followed by a taboo pattern of n bits. The length l of the code is independent of n and can even be zero. Thus, the total length of such a code is l+n for l . 0. The taboo pattern should not appear inside the l code bits and there should be a separator bit at the end of the l-bit code part, to indicate to the decoder the start of the taboo. In the block-based taboo codes, any bit pattern could serve as the taboo, but for the unconstrained codes we have to choose a bit pattern that will prevent ambiguous decoding. The two ideal patterns for the taboo are therefore all zeros or all 1’s, because a string of all zeros can be separated from the preceding l bits by a single 1, and similarly for a string of all 1’s. Other strings require longer separators. Thus, the use of all zeros or 1’s as the taboo string leads to the maximum number of l-bit valid code strings and we will therefore assume that the taboo pattern consists of n consecutive zeros. As a result, an unconstrained taboo code starts with l bits (where l . 0) that do not include n consecutive zeros and that end with a separator bit of 1 (except when l = 0, where the code part consists of no bits and there is no need for a separator). The case n = 1 is special. In this case, the taboo is a single zero, the l bits preceding it cannot include any zeros, so the unconstrained code reduces to the unary code of l 1’s followed by a zero. In order to figure out how to encode and decode these codes, we first have to determine the number of valid bit patterns of l bits, i.e., patterns that do not include any n consecutive zeros. We distinguish three cases. 1. The case l = 0 is trivial. The number of bit patterns of length zero is one, the pattern of no bits. 2. The case 0 < l < n is simple. The string of l bits is too short to have any n consecutive zeros. The last bit must be a 1, so the remaining l - 1 bits can have 2l-1 values. 3. When l . n, we determine the number of valid l-bit strings (i.e., l-bit strings that do not contain any substrings of n zeros) recursively. We denote the number of valid strings by  l n and consider the following cases. When a valid l-bit string starts with a 1 followed by l-1 bits, the number of valid strings is l-1 n . When such a string start with a 01, the number of valid strings is l-2 n . We continue in this way until 134 3. Advanced VL Codes we reach strings that start with 00 . . . 0    n-1 1, where the number of valid strings is l-n n . Thus, the number of valid l-bit strings in this case is the sum shown in the third case of Equation (3.5). l n=... 1 ifl = 0, 2l-1 if l < n, n i=1l-i n  otherwise. (3.5) Table 3.33 lists some values of  l n and it is obvious that the second column (the values of  l 2 ) consists of the well-known Fibonacci numbers. Thus,  l 2  = Fl+1. A deeper look at the table shows that  l 3  = F(3) l+1, where F(3) l = F(3) l-1 + F(3) l-2 + F(3) l-3. This column consists of the n-step Fibonacci numbers of order 3 (tribonaccis), a sequence that starts with F(3) 1 = 1, F(3) 2 = 1, and F(3) 3 = 2. Similarly, the 4th column consists of tetrabonacci numbers, and the remaining columns feature the pentanacci, hexanacci, heptanacci, and other n-step “polynacci” numbers.  l 1   l 2   l 3   l 4   l 5   l 6   l 7   l 8   0 n 1 1 1 1 1 1 1 1  1 n 1 1 1 1 1 1 1 1  2 n 1 2 2 2 2 2 2 2  3 n 1 3 4 4 4 4 4 4  4 n 1 5 7 8 8 8 8 8  5 n 1 8 13 15 16 16 16 16  6 n 1 13 24 29 31 32 32 32  7 n 1 21 44 56 61 63 64 64  8 n 1 34 81 108 120 125 127 128  9 n 1 55 149 208 236 248 253 255 10 n  1 89 274 401 464 492 504 509 Table 3.33: The First Few Values of  l n. The n-step Fibonacci numbers of order k are defined recursively by F(k) l = k i=1 F(k)l-i, (3.6) with initial conditions F(k)1 = 1 and F(k)i = 2i-2 for i = 2, 3, . . . , k. (As an aside, the limit limk›. F(k) is a sequence that starts with an infinite number of 1’s, followed by the powers of 2. The interested reader should also look for anti-Fibonacci numbers, generalized Fibonacci numbers, and other sequences and numbers related to Fibonacci.) We therefore conclude that  l n = F(n) l+1, a relation that is exploited by the developer of 3.17 Wang’s Flag Code 135 these codes to prove that the unconstrained taboo codes are universal (more accurately, can be made as close to universal as desired by increasing n). Table 3.34 lists the organization, lengths, number of codes, and code ranges of some unconstrained taboo codes, and Table 3.35 lists some of these codes for n = 3 (the taboo string is shown in boldface). Codes Length # of values Range t n  0 n 0 to  0 n - 1 1: t 1 + n  1 n  0 n to  0 n +  1 n - 1 ... ... l:t l+ n  l n l-1 i=0 i n to -1 +l i=0 i n ... ... Table 3.34: Organization and Features of Unconstrained Taboo Codes. m Code m Code m Code m Code 0 000 7 111000 14 1111000 21 10011000 1 1000 8 0011000 15 00101000 22 10101000 2 01000 9 0101000 16 00111000 23 10111000 3 11000 10 0111000 17 01001000 24 11001000 4 001000 11 1001000 18 01011000 25 11011000 5 011000 12 1011000 19 01101000 26 11101000 6 101000 13 1101000 20 01111000 27 . . . Table 3.35: Unconstrained Taboo Codes for n = 3. And I promise you, right here and now, no subject will ever be taboo. . . except, of course, the subject that was just under discussion. —Quentin Tarantino, Kill Bill, Vol. 1 The main reference [Pigeon 01b] describes the steps for encoding and decoding these codes. 3.17Wang’s Flag Code Similar to the taboo code, the flag code of MuzhongWang [Wang 88] is based on a flag of f zeros appended to the codewords. The name suffix code is perhaps more appropriate, because the flag must be the suffix of a codeword and cannot appear anywhere inside it. However, this type of suffix code is not the opposite of a prefix code. The principle is to scan the positive integer n that is being encoded and append a single 1 to each sequence of f - 1 zeros found in it. This effectively removes any occurrences of substrings of f consecutive zeros. Before this is done, n is reversed, so that its LSB becomes a 1. This guarantees that the flag will be preceded by a 1 and will therefore be easily recognized by the decoder. 136 3. Advanced VL Codes We use the notation 0f for a string of f zeros, select a value f . 2, and look at two examples. Given n = 66 = 10000102 and f = 3, we reverse n to obtain 0100001 and scan it from left to right, appending a 1 to each string of two consecutive zeros. The result is 010010011 to which is appended the flag. Thus, 010010011|000. Given n = 288 = 1001000002, we reverse it to 000001001, scan it and append 001001010011, and append the flag 001001010011|000. It is obvious that such codewords are easy to decode. The decoder knows the value of f and looks for a string of f - 1 zeros. If such a string is followed by a 0, then it is the flag. The flag is removed and the result is reversed. If such a string is followed by a 1, the 1 is removed and the scan continues. This code, as originally proposed byWang, is slow to implement in software because reversing an m-bit string requires m/2 steps where a pair of bits is swapped in each step. A possible improvement is to move the MSB (which is always a 1) to the right end, where it acts as an intercalary bit to separate the code bits from the flag. Reference [Wang 88] shows that this code is universal, and for large values of n its asymptotic efficiency, defined by Wang as the limit . = lim n›. sup log2(n + 1) L(n) , (3.7) (where L(n) is the length of the Wang code of n) approaches 1. The codeword length L(n) increases with n, but not monotonically because it depends on the number of consecutive zeros in n and on the relation of this number to f. Recall that the binary representation of a power of 2 is a single 1 followed by several consecutive zeros. Thus, integers of the form 2k or slightly bigger tend to have many consecutive zeros. As a result, the Wang codewords of 2k and of its immediate successors tend to be longer than the codewords of its immediate predecessors because of the many intercalary 1’s. On the other hand, the binary representation of the integer 2k -1 consists of k consecutive 1’s, so such numbers and their immediate predecessors tend to have few consecutive zeros, which is why their Wang codewords tend to be shorter. We therefore conclude that the lengths of Wang codewords feature jumps at n values that are powers of 2. It is easy to figure out the length of the Wang code for two types of numbers. The integer n = 2k - 1 = 1k has no zeros in its binary representation, so its Wang code consists of the original k bits plus the f-bit flag. Its successor n+1 = 2k = 10k, on the other hand, has k consecutive zeros, so k/f intercalary bits have to be inserted. The length of the codeword is therefore the original k+1 bits, plus the extra k/f bits, plus the flag, for a total of k +1+k/f + f. Thus, when n passes through a value 2k, the codeword length increases by k/f + 1 bits. Reference [Yamamoto and Ochi 91] shows that in the special case where each bit bj of n (except the MSB b0, which is effectively unused by this code) is selected independently with probability P(bj = 0) = P(bj = 1) = 1/2 for j = 1, 2, . . . , M, the average codeword length L(n) depends on M and f in a simple way (compare with Equation (3.9)) L(n) = M +1+f if M . f - 2, M +1+M-f+2 j=1 lj + f, if M . f - 1, 3.18 Yamamoto Flag Code 137 where the quantity lj is defined recursively by lj =... 0, j. 0, 1 2f-1, j= 1, 1 2f-1 1 2 + lj-(f-1), 2 . j . M - f + 2. (3.8) 3.18 Yamamoto Flag Code Wang’s code requires reversing the bits of the integer n before it is encoded and reversing it again as the last step in its decoding. This code can be somewhat improved if we simply move the MSB of n (which is always a 1) to its right end, to separate the codeword from the flag. The fact that the MSB of n is included in the codeword of n introduces a slight redundancy because this bit is known to be a 1. This bit is needed in Wang’s code, because it helps to identify the flag at the end of a codeword. The flag code of this section, due to Hirosuke Yamamoto and Hiroshi Ochi [Yamamoto and Ochi 91], is more complex to understand and implement, but is faster to encode and decode and does not include the MSB of n. It is therefore slightly shorter on average. In addition, this code is a true flag code, because it is UD even though the bit pattern of the flag may appear inside a codeword. We first explain the encoding process in the simple case f = 3 (a 3-bit flag). We start with a positive integer n = b0b1b2 . . . bM and the particular 3-bit flag p = p1p2p3 = 100. We ignore b0 because it is a 1. Starting at b1, we compare overlapping pairs of bits bibi+1 to the fixed pair p1p2 = 10. If bibi+1 = 10, we append bi to the codeword-so-far and continue with the pair bi+1bi+2. If, however, bibi+1 = 10, we append the triplet 101 (which is bibi+1pf ) to the codeword-so-far and continue with the pair bi+2bi+3. Notice that bibi+1pf = 101 is different from the flag, a fact exploited by the decoder. At the end, when all the pairs bibi+1 have been checked, the LSB bM of n may be left over. This bit is appended to the codeword-so-far, and is followed by the flag. We encode n = 325 = 1010001012 as an example. Its nine bits are numbered b0b1 . . . b8 with b0 = 1. The codeword-so-far starts as an empty string. The first pair is b1b2 = 01 = 10, so b1 = 0 is appended to the codeword-so-far. The next pair is b2b3 = 10, so the triplet 101 is appended to the codeword-so-far, which becomes 0|101. The next pair is b4b5 = 00 = 10, so b4 = 0 is appended to the codeword-so-far. The next pair is b5b6 = 01 = 10, so b5 = 0 is appended to the codeword-so-far, which becomes 0|101|0|0. The next pair is b6b7 = 10, so the triplet 101 is appended to the codeword-so-far, which becomes 0|101|0|0|101. Once b6 and b7 have been included in the codeword-so-far, only b8 = 1 remains and it is simply appended, followed by the flag bits. The resulting codeword is 0|101|0|0|101|1|100 where the two bits in boldface are the complement of p3. Notice that the flag pattern 100 also appears inside the codeword. It is now easy to see how such a codeword can be decoded. The decoder initializes the decoded bitstring to 1 (the MSB of n). Given a string of bits a1a2 . . . whose length is unknown, the decoder scans it, examining overlapping triplets of bits and looking for the pattern aiai+1pf = 101. When such a triplet is found, its last bit is discarded and 138 3. Advanced VL Codes the first two bits 10 are appended to the decoded string. The process stops when the flag is located. In our example, the decoder appends a1 = 0 to the decoded string. It then locates a2a3a4 = 101, discards a4, and appends a3a4 to the decoded string. Notice that the codeword contains the flag pattern 100 in bits a4a5a6 but this pattern disappears once a4 has been discarded. The only potential problem in decoding is that an f-bit pattern of the form . . . aj-1aj |p1p2 . . . (i.e., some bits from the end of the codeword, followed by some bits from the start of the flag) will equal the flag. This problem is solved by selecting the special flag 100. For the special case f = 3, it is easy to verify that bit patterns of the form xx|1 or x|10 cannot equal 100. In the general case, a flag of the form 10f-1 (a 1 followed by f - 1 zeros) is selected, and again it is easy to see that no f-bit string of the form xx. . . x|100 . . . 0 can equal the flag. We are now ready to present the general case, where f can have any value greater than 1. Given a positive integer n = b0b1b2 . . . bM and the f-bit flag p = p1p2 . . . pf = 10f-1, the encoder initializes the codeword-so-far C to the empty string and sets a counter t to 1. It then performs the following loop: 1. If t > M - f + 2 then [if t . M, then C ‹ C + btbt+1 . . . bM endif], C ‹ C + p1p2 . . . pf, Stop. endif 2. If btbt+1 · · · bt+(f-2) = p1p2 . . . pf-1 then [C ‹ C + bt, t ‹ t + 1], else [C ‹ C + btbt+1 . . . bt+(f-2)pf, t ‹ t + (f - 1)] endif Go to step 1. Step 1 should read: If no more tuples remain [if some bits remain, append them to C, endif], append the flag to C. Stop. Step 2 should read: If a tuple (of f - 1 bits from n) does not equal the mostsignificant f -1 bits of the flag, append the next bit bt to C and increment t by 1. Else, append the entire tuple to C, followed by the complement of pf , and increment t by f - 1. Endif. Go to step 1. The developers of this code prove that the choice of the bit pattern 10f-1, or equivalently 01f-1, for the flag guarantees that no string of the form xx. . . x|p1p2 . . . pj can equal the flag. They also propose flags of the form 120f-2 for cases where f . 4. The decoder initializes the decoded string D to 1 (the MSB of n) and a counter s to 1. It then iterates the following step until it finds the flag and stops. If asas+1 . . . as+(f-2) = p1p2 . . . pf-1 then D ‹ D + as, s ‹ s + 1, else [if as+(f-1) = pf then Stop, else D ‹ D + asas+1 . . . as+(f-2), s ‹ s + f endif] endif 3.18 Yamamoto Flag Code 139 The developers also show that this code is universal and is almost asymptotically optimal in the sense of Equation (3.7). The codeword length L(n) increases with n, but not monotonically, and is bounded by log2 n + f . L(n) . log2 n + log2 n f - 1 + f . f f - 1 log2 n + f. In the special case where each bit bj of n (except the MSB b0, which is not used by this code) is selected independently with probability P(bj = 0) = P(bj = 1) = 1/2 for j = 1, 2, . . . , M, the average codeword length L(n) depends on M and f in a simple way (compare with Equation (3.8)) L(n) = M + f if M . f - 2, M + M-f+2 2f-1 + f, if M . f - 1. (3.9) Table 3.36 lists some of the Yamamoto codes for f = 3 and f = 4. These are compared with the similar S(r+1, 01r) codes of Capocelli (Section 3.21.1 and Table 3.41). Recall that the encoder inserts the intercalary bit pf whenever it finds the pattern p1p2 . . . pf-1 in the integer n that is being encoded. Thus, if f is small (a short flag), large values of n may be encoded into long codewords because of the many intercalary bits. On the other hand, large f (a long flag) results in long codewords for small values of n because the flag has to be appended to each codeword. This is why a scheme where the flag starts small and becomes longer with increasing n seems ideal. Such a scheme, dubbed dynamically-variable-flag-length (DVFL), has been proposed by Yamamoto and Ochi as an extension of their original code. The idea is to start with an initial flag length f0 and increment it by 1 at certain points. A function T(f) also has to be chosen that satisfies T(f0) . 1 and T(f + 1) - T(f) . f - 1 for f . f0. Given a large integer n = b0b1 . . . bM, the encoder (and also the decoder, working in lockstep) will increment f when it reaches bits whose indexes equal T(f0), T(f0 + 1), T(f0 + 2), and so on. Thus, bits b1b2 . . . bT(f0) of n will be encoded with a flag length f0, bits bT(f0)+1bT(f0)+2 . . . bT(f0+1) will be encoded with a flag length f0 + 1, and so on. In the original version, the encoder maintains a counter t and examines the f - 1 bits starting at bit bt. These are compared to the first f - 1 bits 10f-2 of the flag. In the extended version, f is determined by the counter t and the function T by solving the inequality T(f - 1) < t . T(f). In the last step, the flag 10fM-1 is appended to the codeword, where fM is determined by the inequality T(fM - 1) M - f + 2 then [if t . M then C ‹ C + btbt+1 . . . bM endif], [if M + 1 > T(f) then f ‹ f + 1 endif], C ‹ C + 10f-1, Stop. endif. 3. If btbt+1 . . . bt+(f-2) = 10f-2 140 3. Advanced VL Codes n S(3, 011) Y3(n) S(4, 0111) Y4(n) 1 011 011 0111 0111 2 0 011 0 011 0 0111 0 0111 3 1 011 1 011 1 0111 1 0111 4 00 011 00 011 00 0111 00 0111 5 01 011 010 011 01 0111 01 0111 6 10 011 10 011 10 0111 10 0111 7 11 011 11 011 11 0111 11 0111 8 000 011 000 011 000 0111 000 0111 9 001 011 0010 011 001 0111 001 0111 10 010 011 0100 011 010 0111 010 0111 11 100 011 0101 011 011 0111 0110 0111 12 101 011 100 011 100 0111 100 0111 13 110 011 1010 011 101 0111 101 0111 14 111 011 110 011 110 0111 110 0111 15 0000 011 111 011 111 0111 111 0111 16 0001 011 0000 011 0000 0111 0000 0111 17 0010 011 00010 011 0001 0111 0001 0111 18 0100 011 00100 011 0010 0111 0010 0111 19 0101 011 00101 011 0011 0111 00110 0111 20 1000 011 01000 011 0100 0111 0100 0111 21 1001 011 010010 011 0101 0111 0101 0111 22 1010 011 01010 011 0110 0111 01100 0111 23 1100 011 01011 011 1000 0111 01101 0111 24 1101 011 1000 011 1001 0111 1000 0111 25 1110 011 10010 011 1010 0111 1001 0111 26 1111 011 10100 011 1011 0111 1010 0111 27 00000 011 10101 011 1100 0111 10110 0111 28 00001 011 1100 011 1101 0111 1100 0111 29 00010 011 11010 011 1110 0111 1101 0111 30 00100 011 1110 011 1111 0111 1110 0111 31 00101 011 1111 011 00000 0111 1111 0111 32 01000 011 00000 011 00001 0111 00000 0111 Table 3.36: Some Capocelli and Yamamoto Codes for f = 3 and f = 4. then C ‹ C + bt, t ‹ t + 1, [if t > T(f) then f ‹ f + 1 endif] else C ‹ C + btbt+1 . . . bt+(f-2)1, t ‹ t + (f - 1), [if t > T(f) then f ‹ f + 1 endif] endif, Go to step 2. The decoding steps are the following: 1. Initialize f ‹ f0, t ‹ 1, s ‹ 1, D ‹ 1. 3.19 Number Bases 141 2. If asas+1 . . . as+(f-2) = 10f-2 then D ‹ D + as, s ‹ s + 1, t ‹ t + 1, [if t > T(f) then f ‹ f + 1 endif], else [if as+(f-1) = 0 then Stop else D ‹ D + asas+1 . . . as+(f-2), s ‹ s + f, t ‹ t + (f - 1), [if t > T(f) then f ‹ f + 1 endif] endif] endif, Go to step 2. The choice of function T(f) is the next issue. The authors show that a function of the form T(f) = Kf(f - 1), where K is a nonnegative constant, results in an asymptotically optimal DVFL. There are other functions that guarantee the same property of DVFL, but the point is that varying f, while generating short codewords, also eliminates one of the chief advantages of any flag code, namely its robustness. If the length f of the flag is increased during the construction of a codeword, then any future error may cause the decoder to lose synchronization and may propagate indefinitely through the string of codewords. It seems that, in practice, the increased reliability achieved by synchronization with a fixed f overshadows the savings produced by a dynamic encoding scheme that varies f. 3.19 Number Bases This short section is a digression to prepare the reader for the Fibonacci and Goldbach codes that follow. Decimal numbers use base 10. The number 203710, for example, has a value of 2 Χ 103 + 0 Χ 102 + 3 Χ 101 + 7 Χ 100. We can say that 2037 is the sum of the digits 2, 0, 3, and 7, each weighted by a power of 10. Fractions are represented in the same way, using negative powers of 10. Thus, 0.82 = 8 Χ 10-1 + 2 Χ 10-2 and 300.7 = 3 Χ 102 + 7 Χ 10-1. Binary numbers use base 2. Such a number is represented as a sum of its digits, each weighted by a power of 2. Thus, 101.112 = 1Χ22+0Χ21+1Χ20+1Χ2-1+1Χ2-2. Since there is nothing special about 10 or 2 (actually there is, because 2 is the smallest integer that can be a base for a number system and 10 is the number of our fingers), it should be easy to convince ourselves that any positive integer n > 1 can serve as the basis for representing numbers. Such a representation requires n “digits” (if n > 10, we use the ten digits and the letters A, B, C, . . . ) and represents the number d3d2d1d0.d-1 as the sum of the digits di, each multiplied by a power of n, thus d3n3 + d2n2 +d1n1 +d0n0 +d-1n-1. The base of a number system does not have to consist of powers of an integer but can be any superadditive sequence that starts with 1. Definition: A superadditive sequence a0, a1, a2, . . . is one where any element ai is greater than the sum of all its predecessors. An example is 1, 2, 4, 8, 16, 32, 64, . . . where each element equals 1 plus the sum of all its predecessors. This sequence consists of the familiar powers of 2, so we know that any integer can be expressed by it using just the digits 0 and 1 (the two bits). Another example is 1, 3, 6, 12, 24, 50, . . . , where each element equals 2 plus the sum of all its predecessors. It is easy to see that any integer can be expressed by it using just the digits 0, 1, and 2 (the three trits). 142 3. Advanced VL Codes Given a positive integer k, the sequence 1, 1 + k, 2+2k, 4+4k, . . . , 2i(1 + k) is superadditive, because each element equals the sum of all its predecessors plus k. Any nonnegative integer can be represented uniquely in such a system as a number x . . . xxy, where x are bits and y is a single digit in the range [0, k]. In contrast, a general superadditive sequence, such as 1, 8, 50, 3102 can be used to represent integers, but not uniquely. The number 50, e.g., equals 8 Χ 6 + 1 + 1, so it can be represented as 0062 = 0 Χ 3102 + 0 Χ 50 + 6 Χ 8 + 2 Χ 1, but also as 0100 = 0 Χ 3102 + 1 Χ 50 + 0 Χ 8 + 0 Χ 1. It can be shown that 1 + r + r2 + · · · + rk is less than rk+1 for any real number r > 1. This implies that the powers of any real number r > 1 can serve as the base of a number system using the digits 0, 1, 2, . . . , d for some d < r. The number . = 1 2 (1 +.5) . 1.618 is the well-known golden ratio. It can serve as the base of a number system, with 0 and 1 as the digits. Thus, for example, 100.1. = .2 + .-1 . 3.2310. Some real bases have special properties. For example, any positive integer R can be expressed as R = b1F1 + b2F2 + b3F3 + b4F5 + b5F8 + b6F13 · · ·, where bi are either 0 or 1, and Fi are the Fibonacci numbers 1, 2, 3, 5, 8, 13, . . . . This representation has the interesting property, known as Zeckendorf’s theorem [Zeckendorf 72], that the string b1b2 . . . does not contain any adjacent 1’s. This useful property, which is the basis for the Goldbach codes (Section 3.22) is easy to prove. If an integer I in this representation has the form . . . 01100 . . ., then because of the definition of the Fibonacci numbers, I can be written . . . 00010 . . .. Examples are the integer 5, whose Fibonacci representation is 0001 and 33 = 1 + 3 + 8 + 21, which is expressed in the Fibonacci base as the 7-bit number 1010101. Section 3.16 discusses the n-step Fibonacci numbers, defined by Equation (3.6). The Australian Aboriginals use a number of languages, some of which employ binary or binary-like counting systems. For example, in the Kala Lagaw Ya language, the numbers 1 through 6 are urapon, ukasar, ukasar-urapon, ukasar-ukasar, ukasar-ukasarurapon, and ukasar-ukasar-ukasar. The familiar terms “dozen” (12) and “gross” (twelve dozen) originated in old duodecimal (base 12) systems of measurement. Computers use binary numbers mostly because it is easy to design electronic circuits with just two states, as opposed to ten states. Leonardo Pisano Fibonacci (1170–1250) Leonard of Pisa (or Fibonacci), was an Italian mathematician, often considered the greatest mathematician of the Middle Ages. He played an important role in reviving ancient mathematics and also made significant original contributions. His book, Liber Abaci, introduced the modern decimal system and the use of Arabic numerals and the zero into Europe. Leonardo was born in Pisa around 1170 or 1180 (he died in 1250). His father Guglielmo was nicknamed Bonaccio (the good natured or simple), which is why Leonardo is known today by the nickname Fibonacci (derived from filius Bonacci, son of Bonaccio). He is also known as Leonardo Pisano, Leonardo Bigollo, Leonardi Bigolli Pisani, Leonardo Bonacci, and Leonardo Fibonacci. 3.20 Fibonacci Code 143 The father was a merchant (and perhaps also the consul for Pisa) in Bugia, a port east of Algiers (now Beja¨ύa, Algeria), and Leonardo visited him there while still a boy. It seems that this was where he learned about the Arabic numeral system. Realizing that representing numbers with Arabic numerals rather than with Roman numerals is easier and greatly simplifies the arithmetic operations, Fibonacci traveled throughout the Mediterranean region to study under the leading Arab mathematicians of the time. Around 1200 he returned to Pisa, where in 1202, at age 32, he published what he had learned in Liber Abaci, or Book of Calculation. Leonardo became a guest of the Emperor Frederick II, who enjoyed mathematics and science. In 1240, the Republic of Pisa honored Leonardo, under the name Leonardo Bigollo, by granting him a salary. Today, the Fibonacci sequence is one of the best known mathematical objects. This sequence and its connection to the golden ratio . = 1 2 (1 + .5) . 1.618 have been studied extensively and the mathematics journal Fibonacci Quarterly is dedicated to the Fibonacci and similar sequences. The publication [Grimm 73] is a short biography of Fibonacci. 3.20 Fibonacci Code The Fibonacci code, as its name suggests, is closely related to the Fibonacci representation of the integers. The Fibonacci code of the positive integer n is the Fibonacci representation of n with an additional 1 appended to the right end. Thus, the Fibonacci code of 5 is 0001|1 and that of 33 is 1010101|1. It is obvious that such a code ends with a pair 11, and that this is the only such pair in the codeword (because the Fibonacci representation does not have adjacent 1’s). This property makes it possible to decode such a code unambiguously, but also causes these codes to be long, because not having adjacent 1’s restricts the number of possible binary patterns. Table 3.37 lists the first 12 Fibonacci codes. 1 11 7 01011 2 011 8 000011 3 0011 9 100011 4 1011 10 010011 5 00011 11 001011 6 10011 12 101011 Table 3.37: Twelve Fibonacci Codes. Decoding. Skip bits of the code until a pair 11 is reached. Replace this 11 by 1. Multiply the skipped bits by the values . . . , 13, 8, 5, 3, 2, 1 (the Fibonacci numbers), 144 3. Advanced VL Codes and add the products. Obviously, it is not necessary to do any multiplication. Simply use the 1 bits to select the proper Fibonacci numbers and add. The Fibonacci codes are long, but have the advantage of being more robust than most other variable-length codes. A corrupt bit in a Fibonacci code may change a pair of consecutive bits from 01 or 10 to 11 or from 11 to 10 or 01. In the former case, a code may be read as two codes, while in the latter case two codes will be decoded as a single code. In either case, the slippage will be confined to just one or two codewords and will not propagate further. The length of the Fibonacci code for n is less than or equal to 1+log. .5n where . is the golden ratio (compare with Equation (2.1)). Figure 3.38 represents the lengths of two Elias codes and the Fibonacci code graphically and compares them to the length of the standard binary (beta) code. 10 100 1000 5 10 15 20 Gamma Length Fib Delta Binary n (* Plot the lengths of four codes 1. staircase plots of binary representation *) bin[i_] := 1 + Floor[Log[2, i]]; Table[{Log[10, n], bin[n]}, {n, 1, 1000, 5}]; g1 = ListPlot[%, AxesOrigin -> {0, 0}, PlotJoined -> True, PlotStyle -> { AbsoluteDashing[{5, 5}]}] (* 2. staircase plot of Fibonacci code length *) fib[i_] := 1 + Floor[ Log[1.618, Sqrt[5] i]]; Table[{Log[10, n], fib[n]}, {n, 1, 1000, 5}]; g2 = ListPlot[%, AxesOrigin -> {0, 0}, PlotJoined -> True] (* 3. staircase plot of gamma code length*) gam[i_] := 1 + 2Floor[Log[2, i]]; Table[{Log[10, n], gam[n]}, {n, 1, 1000, 5}]; g3 = ListPlot[%, AxesOrigin -> {0, 0}, PlotJoined -> True, PlotStyle -> { AbsoluteDashing[{2, 2}]}] (* 4. staircase plot of delta code length*) del[i_] := 1 + Floor[Log[2, i]] + 2Floor[Log[2, Log[2, i]]]; Table[{Log[10, n], del[n]}, {n, 2, 1000, 5}]; g4 = ListPlot[%, AxesOrigin -> {0, 0}, PlotJoined -> True, PlotStyle -> { AbsoluteDashing[{6, 2}]}] Show[g1, g2, g3, g4, PlotRange -> {{0, 3}, {0, 20}}] Figure 3.38: Lengths of Binary, Fibonacci and Two Elias Codes. 3.20 Fibonacci Code 145 In [Fraenkel and Klein 96] (and also [Fraenkel and Klein 85a]), the authors denote this code by C1 and show that it is universal, with c1 = 2 and c2 = 3, but is not asymptotically optimal because c1 > 1. They also prove that for any length r . 1, there are Fr codewords of length r + 1 in C1. As a result, the total number of codewords of length up to and including r is r-1 i=1 Fi = Fr+1 - 1. (See also Figure 3.45.) A P-code is a set of codewords that end with the same pattern P (the pattern is the suffix of the codewords) and where no codeword includes the pattern anywhere else. Given a k-bit binary pattern P, the set of all binary strings of length . k in which P occurs only as a suffix is called the set generated by P and is denoted by L(P). In [Berstel and Perrin 85] this set is called a semaphore code. All codewords in the C1 code end with the pattern P = 11, so this code is L(11). The next Fibonacci code proposed by Fraenkel and Klein is denoted by C2 and is constructed from C1 as follows: 1. Each codeword in C1 ends with two consecutive 1’s; delete one of them. 2. Delete all the codewords that start with 0. Thus, the first few C2 codewords, constructed with the help of Table 3.37, are 1, 101, 1001, 10001, 10101, 100001, and 101001. An equivalent procedure to construct this code is the following: 1. Delete the rightmost 1 of every codeword in C1. 2. Prepend 10 to every codeword. 3. Include 1 as the first codeword. A simple check verifies the equivalence of the two constructions. Code C2 is not a straightforward Fibonacci code as is C1, but it can be termed a Fibonacci code, because the interior bits of each codeword correspond to Fibonacci numbers. The code consists of one codeword of length 1, no codewords of length 2, and Fr-2 codewords of length r for any r . 3. The C2 code is not a prefix code, but is UD. Individual codewords are identified by the decoder because each starts and ends with a 1. Thus, two consecutive 1’s indicate the boundary between codewords. The first codeword introduces a slight complication, but can be handled by the decoder. A string of the form . . . 011110 . . . is interpreted by the decoder as . . . 01|1|1|10 . . ., i.e., two consecutive occurrences of the codeword 1. The C2 code is also more robust than C1. A single error cannot propagate far because the decoder is looking for the pattern 11. The worst case is a string of the form . . . xyz . . . = . . . 1|10 . . . 01|1|10 . . . 01|1 . . . where the middle 1 gets corrupted to a 0. This results in . . . 1|10 . . . 01010 . . . 01|1 . . . which is interpreted by the decoder as one long codeword. The three original codewords xyz are lost, but the error does not propagate any further. Other single errors (corrupted or lost bits) result in the loss of only two codewords. The third Fibonacci code described in [Fraenkel and Klein 96] is denoted by C3 and is constructed from C1 as follows: 1. Delete the rightmost 1 of every codeword of C1. 2. For every r . 1, select the set of C1 codewords of length r, duplicate the set, and distinguish between the two copies by prepending a 10 to all the codewords in one copy and a 11 to all the codewords in the other copy. 146 3. Advanced VL Codes This results in the codewords 101, 111, 1001, 1101, 10001, 10101, 11001, 11101, 100001, 101001, 100101, 110001, . . . . It is easy to see that every codeword of C3 starts with a 1, has at most three consecutive 1’s (and then they appear in the prefix), and every codeword except the second ends with 01. The authors show that for any r . 3 there are 2Fr-2 codewords. It is easy to see that C3 is not a prefix code because, for example, codeword 111 is also the prefix of 11101. However, the code is UD. The decoder first checks for the pattern 111 and interprets it depending on the bit that follows. If that bit is 0, then this is a codeword that starts with 111; otherwise, this is the codeword 111 followed by another codeword. If the current pattern is not 111, the decoder checks for 011. Every codeword except 111 ends with 01, so the pattern 01|1 indicates the end of a codeword and the start of the next one. This pattern does not appear inside any codeword. Given an r-bit codeword y1y2 . . . yr (where yr = 1), the developers show that its index (i.e., the integer whose codeword it is) is given by 2Fr-1 - 2 + y2Fr-2 + r i=3 yiFi-1 - Fr-1 + 1 = r+2 i=3 yiFi-1 + (y2 - 1)Fr-2 - 1, where yr+1 = 1 is the leftmost bit of the next codeword. The developers compare the three Fibonacci codes with the Huffman code using English text of 100,000 words collected from many different sources. Letter distributions were computed and were used to assign Huffman, C1, C2, and C3 codes to the 26 letters. The average lengths of the codes were 4.185, 4.895, 5.298, and 4.891 bits/character, respectively. The Huffman code has the shortest length, but is vulnerable to storage and transmission errors. The conclusion is that the three Fibonacci codes are good candidates for compressing data in applications where robustness is more important than optimal compression. It is interesting to note that the Fibonacci sequence can be generalized by adding a parameter. An m-step Fibonacci number equals the sum of its m immediate predecessors (Equation (3.6)). Thus, the common Fibonacci numbers are 2-step. The m-step Fibonacci numbers can be used to construct order-m Fibonacci codes (Section 3.21), but none of these codes is asymptotically optimal. Section 11.5.4 discusses an application of this type of variable-length code to the compression of sparse strings. The anti-Fibonacci numbers are defined recursively as f(1) = 1, f(2) = 0, and f(k + 2) = f(k) - f(k + 1). The sequence starts with 1, 0, 1, -1, 2, -3, 5, -8, . . . . This sequence is also obtained when the Fibonacci sequence is extended backward from 0. Thus, . . . - 8, 5,-3, 2,-1, 1, 0, 1, 1, 2, 3, 5, 8, . . . . In much the same way that the ratios of successive Fibonacci numbers converge to ., the ratios of successive anti-Fibonacci numbers converge to 1/.. 3.21 Generalized Fibonacci Codes 147 3.21 Generalized Fibonacci Codes The Fibonacci code of Section 3.20 is elegant and robust. The generalized Fibonacci codes presented here, due to Alberto Apostolico and Aviezri Fraenkel [Apostolico and Fraenkel 87], are also elegant, robust, and UD. They are based on m-step Fibonacci numbers (sometimes also called generalized Fibonacci numbers). The authors show that these codes are easy to encode and decode and can be employed to code integers as well as arbitrary, unbound bit strings. (The MSB of an integer is 1, but the MSB of an arbitrary string can be any bit.) The sequence F(m) of m-step Fibonacci numbers F(m) n is defined as follows: F(m) n = F(m) n-1 + F(m) n-2 + · · · + F(m) n-m, for n . 1, and F(m) -m+1 = F(m) -m+2 = · · · = F(m) -2 = 0, F(m) -1 = F(m) 0 = 1. Thus, for example, F(m) 1 = 2 for any m, while for m = 3 we have F(3) -2 = 0, F(3) -1 = F(3) 0 = 1, which implies F(3) 1 = F(3) 0 + F(3) -1 + F(3) -2 = 2, F(3) 2 = F(3) 1 + F(3) 0 + F(3) -1 = 4, and F(3) 3 = F(3) 2 + F(3) 1 + F(3) 0 = 7. The generalized Fibonacci numbers can serve as the basis of a numbering system. Any positive integer N can be represented as the sum of several distinctm-step Fibonacci numbers and this sum does not contain m consecutive 1’s. Thus, if we represent N in this number basis, the representation will not have a run of m consecutive 1’s. An obvious conclusion is that such a run can serve as a comma, to separate consecutive codewords. The two generalized Fibonacci codes proposed by Apostolico and Fraenkel are pattern codes (P-codes) and are also universal and complete (a UD code is complete if the addition of any codeword turns it into a non-UD code). A P-code is a set of codewords that end with the same pattern P (the pattern is the suffix of the codewords) and where no codeword includes the pattern anywhere else. For example, if P = 111, then a P-code is a set of codewords that all end with 111 and none has 111 anywhere else. A P-code is a prefix code because once a codeword c = x . . . x111 has been selected, no other codeword will start with c because no other codeword has 111 other than as a suffix. Given a P-code with the general pattern P = a1a2 . . . ap and an arbitrary codeword x1x2 . . . xna1a2 . . . ap, we consider the string a2a3 . . . ap|x1x2 . . . xna1a2 . . . ap-1 (the end of P followed by the start of a codeword). If P happens to appear anywhere in this string, then the code is not synchronous and a transmission error may propagate over many codewords. On the other hand, if P does not appear anywhere inside such a string, the code is synchronous and is referred to as an SP-code (it is also comma-free). Such codes are useful in applications where data integrity is important. When an error occurs during storage or transmission of an encoded message, the decoder loses synchronization, but regains it when it sees the next occurrence of P (and it sees it when it reads the next codeword). As usual, there is a price to pay for this useful property. In general, SP-codes are not complete. The authors show that the Elias delta code is asymptotically longer than the generalized Fibonacci codes (of any order) for small integers, but becomes shorter at a 148 3. Advanced VL Codes certain point that depends on the order m of the generalized code. For m = 2, this transition point is at F(2) 27 - 1 = 514, 228. For m = 3 it is at (F(3) 63 + F(3) 631 - 1)/2 = 34, 696, 689, 675, 849, 696 . 3.47Χ1016, and for m = 4 it is at (F(4) 231+F(4) 229+F(4) 228-1)/3 . 4.194 Χ 1065. The first generalized Fibonacci code, C(m) 1 , employs the pattern 1m. The code is constructed as follows: 1. The C(m) 1 code of N = 1 is 1m and the C(m) 1 code of N = 2 is 01m. 2. All other C(m) 1 codes have the suffix 01m and a prefix of n-1 bits, where n starts at 2. For each n, there are F(m) n prefixes which are the (n-1)-bit F(m) representations of the integers 0, 1, 2, . . . . Table 3.39 lists examples of C(3) 1 . Once the codes of N = 1 and N = 2 are in place, we set n = 2 and construct the 1-bit F(3) representations of the integers 0, 1, 2, . . . . There are only two such representations, 0 and 1, and they become the prefixes of the next two codewords (for N = 3 and N = 4). We increment n to 3, and construct the 2-bit F(3) representations of the integers 0, 1, 2, . . . . There are four of them, namely 0, 1, 10, and 11. Each is extended to two bits to form 00, 01, 10, and 11, and they become the prefixes of N = 5 through N = 8. String C(3) 1 F(3) M 7421|0111 1 37421 N 0 0 0 111 1 1 00 0111 10 2 000 0|0111 11 3 1 1|0111 100 4 0000 00|0111 101 5 01 01|0111 110 6 2 10| 0111 1000 7 3 11| 0111 1001 8 00000 000|0111 1010 9 001 001|0111 1011 10 02 010|0111 1100 11 03 011|0111 1101 12 4 100|0111 10000 13 5 101|0111 10001 14 6 110|0111 10010 15 000000 0000|0111 10011 16 Table 3.39: The Generalized Fibonacci Code C(3) 1 . C(3) 2 C(2) 2 1 37421|011 1 385321|01 N 11 1 1 1|011 1|01 2 10|011 10|01 3 11|011 100|01 4 100|011 101|01 5 101|011 1000|01 6 110|011 1001|01 7 1000|011 1010|01 8 1001|011 10000|01 9 1010|011 10001|01 10 1011|011 10010|01 11 1100|011 10100|01 12 1101|011 10101|01 13 10000|011 100000|01 14 10001|011 100001|01 15 10010|011 100010|01 16 Table 3.40: Generalized Codes C(3) 2 C(2) 2 . Notice that N1 < N2 implies that the C1 code of N1 is lexicographically smaller than the C1 code of N2 (where a blank space is considered lexicographically smaller than 0). The C1 code is a P-code (with 1m as the pattern) and is therefore a prefix code. 3.21 Generalized Fibonacci Codes 149 The code is partitioned into groups of m+ n codewords each. In each group, the prefix of the codewords is n - 1 bits long and the suffix is 01m. Each group consists of F(m) n-1 codewords of which F(m) n-2 codewords start with 0, F(m) n-3 start with 10, F(m) n-4 start with 110, and so on, up to F(m) n-m-1 codewords that start with 1m-10. A useful, interesting property of this code is that it can be used for arbitrary bitstrings, not just for integers. The difference is that an integer has a MSB of 1, whereas a binary string can start with any bit. The idea is to divide the set of bitstrings into those strings that start with 1 (integers) and those that start with 0. The former strings are assigned C1 codes that start with 1 (in Table 3.39, those are the codes of 4, 7, 8, 13, 14, 15, . . . ). The latter are assigned C1 codes that start with 0 in the following way. Given a bitstring of the form 0iN (where N is a nonnegative integer), assign to it the code 0iN01m. Thus, the string 001 is assigned the code 00|1|0111 (which happens to be the code of 10 in the table). The second generalized Fibonacci code, C(m) 2 , employs the pattern 1m-1. The code is constructed as follows: 1. The C(m) 2 code of N = 1 is 1m-1. 2. All other C(m) 2 codes have the suffix 01m-1 and prefixes that are the F(m) representations of the positive integers, in increasing order. Table 3.40 shows examples of C(3) 2 and C(2) 2 . Once the code of N = 1 has been constructed, we prepare the F(m) representation of n = 1 and prepend it to 01m-1 to obtain the C2 code of N = 2. We then increment n by 1, and construct the other codes in the same way. Notice that C2 is not a prefix code because, for example, C(3) 2 (2) = 1011 is a prefix of C(3) 2 (11) = 1011011. However, C2 is a UD code because the string 01m-1|1 separates any two codewords (each ends with 01m-1 and starts with 1). The C2 codes become longer with N and are organized in length groups for n = 0, 1, 2, . . .. Each group has F(m) n-1-F(m) n-2 codewords of length m+n-1 that can be partitioned as follows: F(m) n-3 codewords that start with 10, F(m) n-4 codewords that start with 110, and so on, up to F(m) n-m-1 codewords that start with 1m-10. The authors provide simple procedures for encoding and decoding the two codes. Once a value for m has been selected, the procedures require tables of many m-step Fibonacci numbers (if even larger numbers are needed, they have to be computed on the fly). A note on robustness. It seems that a P-code is robust. Following an error, the decoder will locate the pattern P very quickly and will resynchronize itself. However, the term “robust” is imprecise and at least code C(3) 1 has a weak point, namely the codeword 111. In his short communication [Capocelli 89], Renato Capocelli points out a case where the decoder of this code can be thrown completely off track because of this codeword. The example is the message 41n3, which is encoded in C(3) 1 as 10111|(111)n|00111. If the second MSB becomes 1 because of an error, the decoder will not sense any error and will decode this string as (111)n+1|1100111, which is the message 1n+1(15). 3.21.1 A Related Code The simple code of this section, proposed by Renato Capocelli [Capocelli 89], is prefix, complete, universal, and also synchronizable (see also Section 4.3 for more synchronous 150 3. Advanced VL Codes codes). It is not a generalized Fibonacci code, but it is related to the C1 and C2 codes of Apostolico and Fraenkel. The code depends on a parameter r and is denoted by S(r + 1, 01r). Once r has been selected, two-part codewords are constructed with a suffix 01r and a prefix that is a binary string that does not contain the suffix. Thus, for example, if r = 2, the suffix is 011 and the prefixes are all the binary strings, starting with the empty string, that do not contain 011. Table 3.41 lists a few examples of S(3, 011) (see also Table 3.36) and it is easy to see how the prefixes are the empty string, 0, 1, 00, 01, and so on, but they include no strings with 011. The codeword of N = 9 is 010|011, but the codeword of N = 10 has the prefix 100 and not 011, so it is 100|011. In general a codeword x . . . x0101y . . . y|011 will be followed by x . . . x1000y . . . y|011 instead of by x . . . x0110y . . . y|011. Such codewords have either the form 0ί|011 (where ί does not contain two consecutive 1’s) or the form 1.|011 (where . does not contain 011). For example, only 12 of the 16 4-bit prefixes can be used by this code, because the four prefixes 0011, 0110, 0111, and 1011 contain the pattern 011. In general, the number of codewords of length N + 3 in S(3, 011) is FN+3 - 1. For N = 4 (codewords of a 4-bit prefix and a 3-bit suffix), the number of codewords is F4+3 -1 = F7 - 1 = 12. N S(3, 011) BS 0 011 0 1 0011 1 2 1011 00 3 00011 01 4 01011 10 5 10011 11 6 11011 000 7 000011 001 8 001011 010 9 010011 011 10 100011 100 11 101011 101 12 110011 110 13 111011 111 Table 3.41: Code S(3, 011) for Integers N and Strings BS. The table also illustrates how S(3, 011) can be employed to encode arbitrary bitstrings, not just integers. Simply write all the binary strings in lexicographic order and assign them the S(3, 011) codewords in increasing order. The special suffix 01r results in synchronized codewords. If the decoder loses synchronization, it regains it as soon as it recognizes the next suffix. What is perhaps more interesting is that the C1 and C2 codes of Apostolico and Fraenkel are special cases of this code. Thus, C(3) 1 is obtained from S(4, 0111) by replacing all the sequences that start with 11 with the sequence 11 and moving the rightmost 1 to the left end of the codeword. Also, C(2) 2 is obtained from S(3, 011) by replacing all the sequences that start with 1 with the codeword 1 and moving the rightmost 1 to the left end of the codeword. The developer provides algorithms for encoding and decoding this code. 3.22 Goldbach Codes 151 Fibonacci couldn’t sleep— Counted rabbits instead of sheep. —Katherine O’Brien 3.22 Goldbach Codes The Fibonacci codes of Section 3.20 have the rare property, termed Zeckendorf’s theorem, that they do not have consecutive 1’s. If a binary number of the form 1xx. . . x does not have any consecutive 1’s, then reversing its bits and appending a 1 results in a number of the form xx. . . x1|1 that ends with two consecutive 1’s, thereby making it easy for a decoder to read individual, variable-length Fibonacci codewords from a long stream of such codes. The variable-length codes described in this section are based on a similar property that stems from the Goldbach conjecture. Goldbach’s conjecture: Every even integer n greater than 2 is the sum of two primes. A prime number is an integer that is not divisible by any other integer (other than trivially by itself and by 1). The integer 2 is prime, but all other primes are odd. Adding two odd numbers results in an even number, so mathematicians have always known that the sum of two primes is even. History had to wait until 1742, when it occurred to Christian Goldbach, an obscure German mathematician, to ask the “opposite” question. If the sum of any two primes is even, is it true that every even integer is the sum of two primes? Goldbach was unable to prove this simple-looking problem, but neither was he able to find a counter-example. He wrote to the famous mathematician Leonhard Euler and received the answer “There is little doubt that this result is true.” However, even Euler was unable to furnish a proof, and at the time of this writing (late 2006), after more than 260 years of research, the Goldbach conjecture has almost, but not quite, been proved. It should be noted that many even numbers can be written as the sum of two primes in several ways. Thus 42 = 23 + 19 = 29 + 13 = 31 + 11 = 37 + 5, 1,000,000 can be partitioned in 5,402 ways, and 100,000,000 has 291,400 Goldbach partitions. Reference [pass 06] is an online calculator that can compute the Goldbach partitions of even integers. Christian Goldbach was a Prussian mathematician. Born in 1690, the son of a pastor, in K¨onigsberg (East Prussia), Goldbach studied law and mathematics. He traveled widely throughout Europe and met with many well-known mathematicians, such as Gottfried Leibniz, Leonhard Euler, and Nicholas (I) Bernoulli. He went to work at the newly opened St Petersburg Academy of Sciences and became tutor to the later Tsar Peter II. The following quotation, from [Mahoney 90] reflects the feelings of his superiors in Russia “. . . a superb command of Latin style and equal fluency in German and French. Goldbach’s polished manners and cosmopolitan circle of friends and acquaintances assured his success in an elite society struggling to emulate its western neighbors.” Goldbach is remembered today for Goldbach’s conjecture. He also studied and proved some theorems on perfect powers. He died in 1764. 152 3. Advanced VL Codes In 2001, Peter Fenwick had the idea of using the Goldbach conjecture (assuming that it is true) to design an entirely new class of codes, based on prime numbers [Fenwick 02]. The prime numbers can serve as the basis of a number system, so if we write an even integer in this system, its representation will have exactly two 1’s. Thus, the even number 20 equals 7 + 13 and can therefore be written 10100, where the five bits are assigned the prime weights (from left to right) 13, 11, 7, 5, and 3. Now reverse this bit pattern so that its least-significant bit becomes 1, to yield 00101. Such a number is easy to read and extract from a long bitstring. Simply stop reading at the second 1. Recall that the unary code (a sequence of zeros terminated by a single 1) is read by a similar rule: stop at the first 1. Thus, the Goldbach codes can be considered an extension of the simple unary code. Goldbach’s original conjecture (sometimes called the “ternary” Goldbach conjecture), written in a June 7, 1742 letter to Euler [dartmouth 06], states “at least it seems that every integer that is greater than 2 is the sum of three primes.” This made sense because Goldbach considered 1 a prime, a convention that is no longer followed. Today, his statement would be rephrased to “every integer greater than 5 is the sum of three primes.” Euler responded with “There is little doubt that this result is true,” and coined the modern form of the conjecture (a form currently referred to as the “strong” or “binary” Goldbach conjecture) which states “all positive even integers greater than 2 can be expressed as the sum of two primes.” Being honest, Euler also added in his response that he regarded this as a fully certain theorem (“ein ganz gewisses Theorema”), even though he was unable to prove it. Reference [utm 06] has information on the history of this and other interesting mathematical conjectures. n 2(n + 3) Primes Codeword 1 8 3+5 11 2 10 3 + 7 101 3 12 5 + 7 011 4 14 3 + 11 1001 5 16 5 + 11 0101 6 18 7 + 11 0011 7 18 7 + 13 00101 8 22 5 + 17 010001 9 24 11 + 13 00011 10 26 7 + 19 0010001 11 28 11 + 17 000101 12 30 13 + 17 000011 13 32 13 + 19 0000101 14 34 11 + 23 00010001 Table 3.42: The Goldbach G0 Code. 3.22 Goldbach Codes 153 The first Goldbach code is designated G0. It encodes the positive integer n by examining the even number 2(n + 3) and writing it as the sum of two primes in reverse (with its most-significant bit at the right end). The G0 codes listed in Table 3.42 are based on the primes (from right to left) 23, 19, 17, 13, 11, 7, 5, and 3. It is obvious that the codes increase in length, but not monotonically, and there is no known expression for the length of G0(n) as a function of n The G0 codes are efficient for small values of n because (1) they are easy to construct based on a table of primes, (2) they are easy to decode, and (3) they are about as long as the Fibonacci codes or the standard binary representation (the ί code) of the integers. However, for large values of n, the G0 codes become too long because, as Table 3.43 illustrates, the primes are denser than the powers of 2 or the Fibonacci numbers. Also, a large even number can normally be written as the sum of two primes in many different ways. For large values of n, a large table of primes is therefore needed and it may take the encoder a while to determine the pair of primes that yields the shortest code for a given large integer n. (The shortest code is normally produced by two primes of similar sizes. Writing n = a + b where a is small and b is large, results in a long G0 code. The best Goldbach partition for 11,230, for example, is 2003 + 9227 and the worst one is 17 + 11213.) n: 1 2 3 4 5 6 7 8 9 10 11 12 Pn: 1 3 5 7 11 13 17 19 23 29 31 37 2n-1: 1 2 4 8 16 32 64 128 256 512 1024 2048 Fn: 1 1 2 3 5 8 13 21 34 55 89 144 Table 3.43: Growth of Primes, Powers of 2, and Fibonacci Numbers. Thus, the G0 code of a large integer n is long, and since it has only two 1’s, it must have many zeros and may have one or two runs of zeros. This property is the basis for the Goldbach G1 code. The principle of G1 is to determine two primes Pi and Pj (where i . j) whose sum yields a given integer n, and encode the pair (i, j -i+1) with two gamma codes. Thus, for example, n = 100 can be written as the following sums 3 + 97, 11 + 89, 17 + 83, 29 + 71, 41 + 59, and 47 + 53. We select 47 + 53 = P15 + P16, yielding the pair (15, 16 - 15 + 1 = 2) and the two gamma codes 0001111:010 that are concatenated to form the G1 code of 100. For comparison, selecting the Goldbach partition 100 = 3 + 97 yields the indexes 2 and 25, the pair (2, 25 - 2 + 1), and the two gamma codes 010:000011000, two bits longer. Notice that i may equal j, which is why the pair (i, j - i + 1) and not (i, j - i) is encoded. The latter may result in a second element of 0. Table 3.44 lists several G1 codes of even integers. It is again obvious that the lengths of the G1 codes increase but not monotonically. The lengths of the corresponding gamma codes are also listed for comparison, and it is clear that the G1 code is the winner in most cases. Figure 3.45 illustrates the lengths of the G1, Elias gamma code, and the C1 code of Fraenkel and Klein (Section 3.20). The last two rows of Table 3.44 list two alternatives for the G1 code of 40, and they really are two bits longer than the best code of 40. However, selecting the Goldbach partition with the most similar primes does not always yield the shortest code, as in the 154 3. Advanced VL Codes n Sum Indexes Pair Codeword Len. |.(n)| 2 1+1 1,1 1,1 1:1 2 3 4 1+3 1,2 1,2 1:010 4 5 6 3+3 2,2 2,1 010:1 4 5 8 3+5 2,3 2,2 010:010 6 7 10 3+7 2,4 2,3 010:011 6 7 12 5+7 3,4 3,2 011:010 6 7 14 7+7 4,4 4,1 00100:1 6 7 16 5+11 3,5 3,3 011:011 6 9 18 7+11 4,5 4,2 00100:011 8 9 20 7+13 4,6 4,3 00100:011 8 9 30 13+17 6,7 6,2 00110:010 8 9 40 17+23 7,9 7,3 00111:011 8 11 50 13+37 6,12 6,7 00110:00111 10 11 60 29+31 10,11 10,2 0001010:010 10 11 70 29+41 10,13 10,4 0001010:00100 12 13 80 37+43 12,14 12,3 0001100:011 10 13 90 43+47 14,15 14,2 0001110:010 10 13 100 47+53 15,16 15,2 0001111:010 10 13 40 3+37 2,12 2,11 010:0001011 10 11 40 11+29 5,10 5,6 00101:00110 10 11 Table 3.44: The Goldbach G1 Code. case of 50 = 13+37 = 19+31. Selecting the first pair (where the two primes differ most) yields the indexes 6, 12 and the pair (6, 7) for a total gamma codes of 10 bits. Selecting the second pair (the most similar primes), on the other hand, results in indexes 8, 11 and a pair (8, 4) for a total of 12 bits of gamma codes. In order for it to be useful, the G1 code has to be extended to arbitrary positive integers, a task that is done in two ways as follows: 1. Map the positive integer n to the even number N = 2(n + 3) and encode N in G1. This is a natural extension of G1, but it results in indexes that are about 60% larger and thus generate long codes. For example, the extended code for 20 would be the original G1 code of 46 = 17 + 29. The indexes are 7, 10, the pair is (7, 4), and the codes are 00111:00100, two bits longer than the original G1 code of 20. 2. A better way to extend G1 is to consider various cases and handle each differently so as to obtain the best codes in every case. The developer, Peter Fenwick, refers to the result as the G2 code. The cases are as follows: 2.1. The integers 1 and 2 are encoded as 110 and 111, respectively (no other codes will start with 11). 2.2. The even integers are encoded as in G1, but with a small difference. Once it is determined that n = Pi +Pj , we encode the pair (i+1, j -i+1) instead of (i, j -i+1). Thus, if i = 1, it is encoded as 010, the gamma code of 2. This guarantees that the G2 code of an even integer will not start with a 1 and will always have the form 0. . . :0. . . . 2.3. If n is the prime Pi, it is encoded as the gamma code of (i + 1) followed by a 3.22 Goldbach Codes 155 4 8 16 32 64 128 256 512 2 4 6 8 10 12 14 16 18 20 Codeword length Goldbach G1 Elias gamma F&K C1 n Clear[g1, g2, g3]; n=Table[2^i, {i,2,9}]; (* ELias gamma *) Table[{i, IntegerPart[1+2Log[2, n[[i]]]]}, {i,1,8}]; g1=ListPlot[%, PlotJoined->True]; (* Goldbach G2 *) m[i_]:=4IntegerPart[Log[2, (1.15n[[i]]/2)/Log[E, n[[i]]/2]]]; Table[{i, m[i]}, {i,1,8}]; g2=ListPlot[%, PlotJoined->True, PlotStyle->{Dashing[{0.05, 0.05}]}]; (* Fraenkel Klein C^1 *) Table[{i, IntegerPart[Log[GoldenRatio, n[[i]] Sqrt[5]]+1]}, {i,1,8}]; g3=ListPlot[%, PlotJoined->True, PlotStyle->{Dashing[{0.01,0.03,0.02,0.03}]}]; Show[g1, g2, g3, PlotRange->{{0, 8},{0, 20}}, Ticks->{{{1,"4"},{2,"8"},{3,"16"},{4,"32"}, {5,"64"},{6,"128"},{7,"256"},{8,"512"}}, {0, 2, 4, 6, 8, 10, 12, 14, 16, 18, 20}}, TextStyle->{FontFamily->"cmr10", FontSize->10}] Figure 3.45: Codeword Length Versus n for Three Codes. single 1 to yield 0. . . :1. 2.4. If n is odd but is not a prime, its G2 code starts with a single 1 followed by the G2 code of the even number n - 1. The resulting gamma codes have the form 1:0. . . :0. . . . Table 3.46 lists some examples of the G2 code and compares their lengths to the lengths of the gamma and omega codes. In most cases, G2 is the shortest of the three, but in some cases, most notably when n is a power of 2, G2 is longer than the other codes. Given an even integer n, there is no known formula to locate any of its Goldbach partitions. Thus, we cannot determine the precise length of any Goldbach code, but we can provide an estimate, and the G1 code length is the easiest to estimate. We assume that the even number 2n is the sum of two primes, each of order n. According to the prime number theorem, the number .(n) of primes less than n is approximately proportional to n/ ln n. Extensive experiments carried out by Peter Fenwick, the developer of 156 3. Advanced VL Codes n Codeword Len. |.(n)| |.(n)| 1 110 3 1 1 2 111 3 3 3 3 0101 4 3 3 4 010010 6 5 6 5 0111 4 5 6 6 011010 6 5 6 7 001001 6 5 6 8 011011 6 7 7 9 1011011 7 7 7 10 01100100 8 7 7 11 001011 6 7 7 12 00101010 8 7 7 13 001101 6 7 7 14 00110010 8 7 7 15 100110010 9 7 7 16 0010100100 10 9 11 17 001111 6 9 11 18 00111010 8 9 11 19 00010001 8 9 11 20 00111011 8 9 11 30 0001010010 10 9 11 40 0001100011 10 11 12 50 0001111011 10 11 12 60 000010001010 12 11 12 70 000010011011 12 13 13 80 000010110010 12 13 13 90 000011000010 12 13 13 100 000011001011 12 13 13 Table 3.46: The Goldbach G2 Code. these codes, indicate that for values up to 1000, a good constant of proportionality is 1.15. Thus, we estimate index i at i . 1.15n/ ln n, resulting in a gamma code of 2 log2 1.15n ln n !+ 1 def = 2L + 1 bits. (3.10) We estimate the second element of the pair (i, j -i+1) at i/2 (Table 3.44 shows that it is normally smaller), which implies that the second gamma code is two bits shorter than the first. The total length of the G1 code is therefore (2L + 1) + (2L - 1) = 4L, where L is given by Equation (3.10). Direct computations show that the G1 code is generally shorter than the gamma code for values up to 100 and longer for larger values. The length of the G2 code is more difficult to estimate, but a simple experiment that computed these codes for n values from 2 to 512 indicates that their lengths (which often vary by 2–3 bits from code to code) can be approximated by the smooth function 3.23 Additive Codes 157 2+ 13 8 log2 n. For n = 512, this expression has the value 16.625, which is 1.66 times the length of the binary representation of 512 (10 bits). This should be contrasted with the gamma code, where the corresponding factor is 2. I can envision an abstract of a paper, circa 2100, that reads: “We can show, in a certain precise sense, that the Goldbach conjecture is true with probability larger than 0.99999, and that its complete truth could be determined with a budget of $10B.”) —Doron Zeilberger (1993) 3.23 Additive Codes The fact that the Fibonacci numbers and the Goldbach conjecture lead to simple, efficient codes suggests that other number sequences may be employed in the same way to create similar, and perhaps even better, codes. Specifically, we may extend the Goldbach codes if we find a sequence S of “basis” integers such that any integer n can be expressed as the sum of two elements of S (it is possible to search for similar sequences that can express any n as the sum of three, four, or more elements, but we restrict our search to sums of two sequence elements). Given such a sequence, we can conceive a code similar to G0, but for all the nonnegative integers, not just the even integers. Given several such sequences, we can call the resulting codes “additive codes.” One technique to generate a sequence of basis integers is based on the sieve principle (compare with the sieve of Eratosthenes). We start with a short sequence (a basis set) whose elements ai are sufficient to represent each of the first few positive integers as a sum ai + aj . This initial sequence is then enlarged in steps. In a general step, we have just added a new element ak and ended up with the sequence S = (0, a1, a2, . . . , ak) such that each integer 1 . n . ak can be represented as a sum of two elements of S. We now generate all the sums ai + ak for i values from 1 to k and check to verify that they include ak + 1, ak + 2, and so on. When we find the first missing integer, we append it to S as element ak+1. We now know that each integer 1 . n . ak+1 can be represented as a sum of two elements from S, and we can continue to extend sequence S of basis integers. This sieve technique is illustrated with the basis set (0, 1, 2). Any of 0, 1, and 2 can be represented as the sum of two elements from this set. We generate all the possible sums ai + ak = ai + 2 and obtain 3 and 4. Thus, 5 is the next integer that cannot be generated, and it becomes the next element of the set. Adding ai +5 yields 5, 6, and 7, so the next element is 8. Adding ai + 8 in the sequence (0, 1, 2, 5, 8) yields 8, 9, 10, 13, and 16, so the next element should be 11. Adding ai+11 in the sequence (0, 1, 2, 5, 8, 11) yields 11, 12, 13, 16, 19, and 22, so the next element should be 14. Continuing in this way, we obtain the sequence 0, 1, 2, 5, 8, 11, 14, 16, 20, 23, 26, 29, 33, 46, 50, 63, 67, 80, 84, 97, 101, 114, 118, 131, 135, 148, 152, 165, 169, 182, 186, 199, 203, 216, 220, 233, and 237 whose 37 elements can represent every integer from 0 to 250 as the sum of two elements. The equivalent Goldbach code for integers up to 250 requires 53 primes, so it has the potential of being longer. The initial basis set may be as small as (0, 1), but may also contain other integers (seeds). In fact, the seeds determine the content and performance of the additive sequence. Starting with just the smallest basic set (0, 1), adding ai + 1 yields 1 and 2, 158 3. Advanced VL Codes so the next element should be 3. Adding ai + 3 in the sequence (0, 1, 3) yields 3, 4, and 6, so the next element should be 5. Adding ai + 5 in the sequence (0, 1, 3, 5) yields 5, 6, 8, and 10, so the next element should be 7. We end up with a sequence of only odd numbers, which is why it may be a good idea to start with the basic sequence plus some even seeds. Table 3.47 lists some additive codes based on the additive sequence (0, 1, 3, 5, 7, 9, 11, 12, 25, 27, 29, 31, 33, 35, . . .) and compares their lengths to the lengths of the corresponding gamma codes. n Sum Indexes Pair Codeword Len. |.(n)| 10 3 + 7 3,5 3,3 011:011 6 7 11 0 + 11 1,7 1,7 1:00111 6 7 12 1 + 11 2,7 2,6 010:00110 8 7 13 1 + 12 2,8 2,7 010:00111 8 7 14 7 + 7 5,5 5,1 00101:1 6 7 15 3 + 12 3,8 3,6 011:00110 8 7 16 7 + 9 5,6 5,2 00101:010 8 9 17 5 + 12 4,8 4,5 00100:00101 10 9 18 7 + 11 5,7 5,3 00101:011 8 9 20 9 + 11 6,7 6,2 00110:010 8 9 30 5 + 25 4,9 4,6 00100:00110 10 9 40 11 + 29 7,11 7,5 00111:00101 10 11 50 25 + 25 9,9 9,1 0001001:1 8 11 60 29 + 31 11,12 11,2 0001011:010 10 11 70 35 + 35 14,14 14,1 0001110:1 8 13 80 0 + 80 1,20 1,20 1:000010100 10 13 90 29 + 61 11,17 11,7 0001011:00111 14 13 100 3 + 97 3,23 3,21 011:000010101 14 13 Table 3.47: An Additive Code. The developer of these codes presents a summary where the gamma, delta, Fibonacci, G1, and additive codes are compared for integers n from 1 to 256, and concludes that for values from 4 to 30, the additive code is the best of the five, and for the entire range, it is better than the gamma and delta codes. Given integer data in a certain interval, it seems that the best additive codes to compress the data are those generated by the shortest additive sequence. Determining this sequence can be done by a brute force approach that tries many sets of seeds and computes the additive sequence generated by each set. Ulam Sequence. The Ulam sequence (u, v) is defined by a1 = u, a2 = v, and ai for i > 2 is the smallest integer that can be expressed uniquely as the sum of two distinct earlier elements. The numbers generated this way are sometimes called u-numbers or Ulam numbers. A basic reference is [Ulam 06]. The first few elements of the (1, 2)-Ulam sequence are 1, 2, 3, 4, 6, 8, 11, 13, 16,. . . . The element following the initial (1, 2) is 3, because 3 = 1 + 2. The next element is 4 = 1 + 3. (There is no need to worry about 4 = 2 + 2, because this is a sum of two 3.23 Additive Codes 159 identical elements instead of two distinct elements.) The integer 5 is not an element of the sequence because it is not uniquely representable, but 6 = 2 + 4 is. These simple rules make it possible to generate Ulam sequences for any pair (u, v) of integers. Table 3.48 lists several examples (the “Sloane” labels in the second column refer to integer sequences from [Sloane 06]). (u, v) Sloane Sequence (1, 2) A002858 1, 2, 3, 4, 6, 8, 11, 13, 16, 18, . . . (1, 3) A002859 1, 3, 4, 5, 6, 8, 10, 12, 17, 21, . . . (1, 4) A003666 1, 4, 5, 6, 7, 8, 10, 16, 18, 19, . . . (1, 5) A003667 1, 5, 6, 7, 8, 9, 10, 12, 20, 22, . . . (2, 3) A001857 2, 3, 5, 7, 8, 9, 13, 14, 18, 19, . . . (2, 4) A048951 2, 4, 6, 8, 12, 16, 22, 26, 32, 36, . . . (2, 5) A007300 2, 5, 7, 9, 11, 12, 13, 15, 19, 23, . . . Table 3.48: Some Ulam Sequences. It is clear that Ulam sequences are additive and can be used to generate additive codes. The following Mathematica code to generate such a sequence is from [Sloane 06], sequence A002858. Ulam4Compiled = Compile[{{nmax, _Integer}, {init, _Integer, 1}, {s, _Integer}}, Module[{ulamhash = Table[0, {nmax}], ulam = init}, ulamhash[[ulam]] = 1; Do[ If[Quotient[Plus @@ ulamhash[[i - ulam]], 2] == s, AppendTo[ulam, i]; ulamhash[[i]] = 1], {i, Last[init] + 1, nmax}]; ulam]]; Ulam4Compiled[355, {1, 2}, 1] Stanislaw Marcin Ulam (1909–1984) One of the most prolific mathematicians of the 20th century, Stanislaw Ulam also had interests in astronomy and physics. He worked on the hydrogen bomb at the Los Alamos National Laboratory, proposed the Orion project for nuclear propulsion of space vehicles, and originated the Monte-Carlo method (and also coined this term). In addition to his important original work, mostly in point set topology, Ulam was also interested in mathematical recreations, games, and oddities. The following quotation, by his friend and colleague Gian-Carlo Rota, summarizes Ulam’s unusual personality and talents. “Ulam’s mind is a repository of thousands of stories, tales, jokes, epigrams, remarks, puzzles, tongue-twisters, footnotes, conclusions, slogans, formulas, diagrams, quotations, limericks, summaries, quips, epitaphs, and headlines. In the course of a normal conversation he simply pulls out of his mind the fifty-odd relevant items, and presents them in linear succession. A second-order memory prevents him from repeating himself too often before the same public.” 160 3. Advanced VL Codes There is another Ulam sequence that is constructed by a simple rule. Start with any positive integer n and construct a sequence as follows: 1. If n = 1, stop. 2. If n is even, the next number is n/2; go to step 1. 3. If n is odd, the next number is 3n + 1; go to step 1. Here are some examples for the first few integers: (1), (10, 5, 16, 8, 4, 2, 1), (2, 1), (16, 8, 4, 2, 1), (3, 10, 5, . . . ) (22, 11, 34, 17, 52, 26, 13, 40, 20, 10, 5, . . . ). These sequences are sometimes known as the 3x + 1 problem, because no one has proved that they always reach the value 1. However, direct checking up to 100 Χ 250 suggests that this may be so. 3.24 Golomb Code The seventeenth-century French mathematician Blaise Pascal is known today mostly for his contributions to the field of probability, but during his short life he made important contributions to many areas. It is generally agreed today that he invented (an early version of) the game of roulette (although some believe that this game originated in China and was brought to Europe by Dominican monks who were trading with the Chinese). The modern version of roulette appeared in 1842. The roulette wheel has 37 shallow depressions (known as slots) numbered 0 through 36 (the American version has 38 slots numbered 00, 0, and 1 through 36). The dealer (croupier) spins the wheel while sending a small ball rolling in the opposite direction inside the wheel. Players can place bets during the spin until the dealer says “no more bets.” When the wheel stops, the slot where the ball landed determines the outcome of the game. Players who bet on the winning number are paid according to the type of bet they placed, while players who bet on the other numbers lose their entire bets to the house. [Bass 92] is an entertaining account of an attempt to compute the result of a roulette spin in real time. The simplest type of bet is on a single number. A player winning this bet is paid 35 times the amount bet. Thus, a player who plays the game repeatedly and bets $1 each time expects to lose 36 games and win one game out of every set of 37 games on average. The player therefore loses on average $37 for every $35 won. The probability of winning a game is p = 1/37 . 0.027027 and that of losing a game is the much higher q = 1-p = 36/37 . 0.972973. The probability P(n) of winning once and losing n - 1 times in a sequence of n games is the product qn-1p. This probability is normalized because . n=1 P(n) = . n=1 qn-1p = p . n=0 qn = p 1 - q = p p = 1. 3.24 Golomb Code 161 As n grows, P(n) shrinks slowly because of the much higher value of q. The values of P(n) for n = 1, 2, . . . , 10 are 0.027027, 0.026297, 0.025586, 0.024895, 0.024222, 0.023567, 0.022930, 0.022310, 0.021707, and 0.021120. The probability function P(n) is said to obey a geometric distribution. The reason for the name “geometric” is the resemblance of this distribution to the geometric sequence. A sequence where the ratio between consecutive elements is a constant q is called geometric. Such a sequence has elements a, aq, aq2, aq3, . . . . The (infinite) sum of these elements is a geometric series . i=0 aqi. The interesting case is where q satisfies -1 < q < 1, in which the series converges to a/(1-q). Figure 3.49 shows the geometric distribution for p = 0.2, 0.5, and 0.8. 2 4 6 8 10 0.2 p=0.2 0.4 0.6 p=0.5 0.8 p=0.8 Figure 3.49: Geometric Distributions for p = 0.2, 0.5, and 0.8. Certain data compression methods are based on run-length encoding (RLE). Imagine a binary string where a 0 appears with probability p and a 1 appears with probability 1-p. If p is large, there will be runs of zeros, suggesting the use of RLE to compress the string. The probability of a run of n zeros is pn, and the probability of a run of n zeros followed by a 1 is pn(1-p), indicating that run lengths are distributed geometrically. A naive approach to compressing such a string is to compute the probability of each run length and apply the Huffman method to obtain the best prefix codes for the run lengths. In practice, however, there may be a large number of run lengths and this number may not be known in advance. A better approach is to construct an infinite family of optimal prefix codes, such that no matter how long a run is, there will be a code in the family to encode it. The codes in the family must depend on the probability p, so we are looking for an infinite family of parametrized prefix codes. The Golomb codes presented here [Golomb 66] are such codes and they are the best ones for the compression of data items that are distributed geometrically. Let’s first examine a few numbers to see why such codes must depend on p. For p = 0.99, the probabilities of runs of two zeros and of 10 zeros are 0.992 = 0.9801 and 0.9910 = 0.9, respectively (both large). In contrast, for p = 0.6, the same run lengths have the much smaller probabilities of 0.36 and 0.006. The ratio 0.9801/0.36 is 2.7225, but the ratio 0.9/0.006 is the much greater 150. Thus, a large p implies higher probabilities for long runs, whereas a small p implies that long runs will be rare. 162 3. Advanced VL Codes Two relevant statistical concepts are the mean and median of a sequence of run lengths. They are illustrated by the binary string 00000100110001010000001110100010000010001001000110100001001 (3.11) that has the 18 run lengths 5, 2, 0, 3, 1, 6, 0, 0, 1, 3, 5, 3, 2, 3, 0, 1, 4, and 2. Its mean is the average (5+2+0+3+1+6+0+0+1+3+5+3+2+3+0+1+4+2)/18 . 2.28. Its median m is the value such that about half the run lengths are shorter than m and about half are equal to or greater than m. To find m, we sort the 18 run lengths to obtain 0, 0, 0, 0, 1, 1, 1, 2, 2, 2, 3, 3, 3, 3, 4, 5, 5, and 6 and find that the median (the central number) is 2. We are now ready for a description of the Golomb code. The main feature of this code is its coding efficiency when the data consists of two asymmetric events, one common and the other one rare, that are interleaved. Encoding. The Golomb code for nonnegative integers n depends on the choice of a parameter m (we’ll see later that for RLE, m should depend on the probability p and on the median of the run lengths). Thus, it is a parametrized prefix code, which makes it especially useful in cases where good values for the parameter can be computed or estimated. The first step in constructing the Golomb code of the nonnegative integer n is to compute the three quantities q (quotient), r (remainder), and c by q = "n m#, r= n - qm, and c = log2m, following which the code is constructed in two parts; the first is the value of q, coded in unary, and the second is the binary value of r coded in a special way. The first 2c - m values of r are coded, as unsigned integers, in c - 1 bits each, and the rest are coded in c bits each (ending with the biggest c-bit number, which consists of c 1’s). The case where m is a power of 2 (m = 2c) is special because it requires no (c-1)-bit codes. We know that n = r + qm; so once a Golomb code is decoded, the values of q and r can be used to easily reconstruct n. The case m = 1 is also special. In this case, q = n and r = c = 0, implying that the Golomb code of n is its unary code. Examples. Choosing m = 3 produces c = 2 and the three remainders 0, 1, and 2. We compute 22 - 3 = 1, so the first remainder is coded in c - 1 = 1 bit to become 0, and the remaining two are coded in two bits each ending with 112, to become 10 and 11. Selecting m = 5 results in c = 3 and produces the five remainders 0 through 4. The first three (23-5 = 3) are coded in c-1 = 2 bits each, and the remaining two are each coded in three bits ending with 1112. Thus, 00, 01, 10, 110, and 111. The following simple rule shows how to encode the c-bit numbers such that the last of them will consist of c 1’s. Denote the largest of the (c - 1)-bit numbers by b, then construct the integer b+1 in c-1 bits, and append a zero on the right. The result is the first of the c-bit numbers and the remaining ones are obtained by incrementing. Table 3.50 shows some examples of m, c, and 2c-m, as well as some Golomb codes for m = 2 through 13. For a somewhat longer example, we select m = 14. This results in c = 4 and produces the 14 remainders 0 through 13. The first two (24 - 14 = 2) are coded in c - 1 = 3 bits each, and the remaining 12 are coded in four bits each, ending with 3.24 Golomb Code 163 m 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 c 1 2 2 3 3 3 3 4 4 4 4 4 4 4 4 2c -m 0 1 0 3 2 1 0 7 6 5 4 3 2 1 0 m/n 0 1 2 3 4 5 6 7 8 9 10 11 12 2 0|0 0|1 10|0 10|1 110|0 110|1 1110|0 1110|1 11110|0 11110|1 111110|0 111110|1 1111110|0 3 0|0 0|10 0|11 10|0 10|10 10|11 110|0 110|10 110|11 1110|0 1110|10 1110|11 11110|0 4 0|00 0|01 0|10 0|11 10|00 10|01 10|10 10|11 110|00 110|01 110|10 110|11 11110|00 5 0|00 0|01 0|10 0|110 0|111 10|00 10|01 10|10 10|110 10|111 110|00 110|01 110|10 6 0|00 0|01 0|100 0|101 0|110 0|111 10|00 10|01 10|100 10|101 10|110 10|111 110|00 7 0|00 0|010 0|011 0|100 0|101 0|110 0|111 10|00 10|010 10|011 10|100 10|101 10|110 8 0|000 0|001 0|010 0|011 0|100 0|101 0|110 0|111 10|000 10|001 10|010 10|011 10|100 9 0|000 0|001 0|010 0|011 0|100 0|101 0|110 0|1110 0|1111 10|000 10|001 10|010 10|011 10 0|000 0|001 0|010 0|011 0|100 0|101 0|1100 0|1101 0|1110 0|1111 10|000 10|001 10|010 11 0|000 0|001 0|010 0|011 0|100 0|1010 0|1011 0|1100 0|1101 0|1110 0|1111 10|000 10|001 12 0|000 0|001 0|010 0|011 0|1000 0|1001 0|1010 0|1011 0|1100 0|1101 0|1110 0|1111 10|000 13 0|000 0|001 0|010 0|0110 0|0111 0|1000 0|1001 0|1010 0|1011 0|1100 0|1101 0|1110 0|1111 Table 3.50: Some Golomb Codes for m = 2 Through 13. 11112 (and as a result starting with 01002). Thus, we have 000, 001, followed by the 12 values 0100, 0101, 0110, 0111, . . . , 1111. Table 3.51 lists several detailed examples and Table 3.52 lists 48 codes for m = 14 and for m = 16. The former starts with two 4-bit codes, followed by sets of 14 codes each that are getting longer by one bit. The latter is simpler because 16 is a power of 2. The Golomb codes for m = 16 consist of sets of 16 codes each that get longer by one bit. The Golomb codes for the case where m is a power of 2 have been conceived by Robert F. Rice and are called Rice codes. They are employed by several algorithms for lossless audio compression. n 0 1 2 3 . . . 13 14 15 16 17 . . . 27 28 29 30 q =  n 14  0 0 0 0 . . . 0 1 1 1 1 . . . 1 2 2 2 unary(q) 0 0 0 0 . . . 0 10 10 10 10 . . . 10 110 110 110 r 000 001 0100 0101 . . . 1111 000 001 0100 0101 . . . 1111 000 001 0100 Table 3.51: Some Golomb Codes for m = 14. Tables 3.51 and 3.52 illustrate the effect of m on the code length. For small values of m, the Golomb codes start short and rapidly increase in length. They are appropriate for RLE in cases where the probability p of a 0 bit is small, implying very few long runs. For large values of m, the initial codes (for n = 1, 2, . . .) are long, but their lengths increase slowly. Such codes make sense for RLE when p is large, implying that many long runs are expected. Decoding. The Golomb codes are designed in this special way to facilitate their decoding. We first demonstrate the decoding for the simple case m = 16 (m is a power of 2). To decode, start at the left end of the code and count the number A of 1’s preceding the first 0. The length of the code is A + c + 1 bits (for m = 16, this is A + 5 bits). If we denote the rightmost five bits of the code by R, then the value of the code is 16A+R. This simple decoding reflects the way the code was constructed. To encode n with m = 16, start by dividing it by 16 to get n = 16A + R, then write A 1’s followed by a single 0, followed by the 4-bit representation of R. 164 3. Advanced VL Codes m = 14 m = 16 n Code n Code n Code n Code 0 0000 24 101100 0 00000 24 101000 1 0001 25 101101 1 00001 25 101001 26 101110 2 00010 26 101010 2 00100 27 101111 3 00011 27 101011 3 00101 28 110000 4 00100 28 101100 4 00110 29 110001 5 00101 29 101101 5 00111 6 00110 30 101110 6 01000 30 1100100 7 00111 31 101111 7 01001 31 1100101 8 01000 8 01010 32 1100110 9 01001 32 1100000 9 01011 33 1100111 10 01010 33 1100001 10 01100 34 1101000 11 01011 34 1100010 11 01101 35 1101001 12 01100 35 1100011 12 01110 36 1101010 13 01101 36 1100100 13 01111 37 1101011 14 01110 37 1100101 14 10000 38 1101100 15 01111 38 1100110 15 10001 39 1101101 39 1100111 40 1101110 16 100000 40 1101000 16 100100 41 1101111 17 100001 41 1101001 17 100101 42 1110000 18 100010 42 1101010 18 100110 43 1110001 19 100011 43 1101011 19 100111 20 100100 44 1101100 20 101000 44 11100100 21 100101 45 1101101 21 101001 45 11100101 22 100110 46 1101110 22 101010 46 11100110 23 100111 47 1101111 23 101011 47 11100111 Table 3.52: The First 48 Golomb Codes for m = 14 and m = 16. For m values that are not powers of 2, decoding is slightly more involved. Assuming again that a code begins with A 1’s, start by removing them and the zero immediately following them. Denote the c - 1 bits that follow by R. If R < 2c - m, then the total length of the code is A + 1 + (c - 1) (the A 1’s, the zero following them, and the c - 1 bits that follow) and its value is mΧA + R. If R . 2c -m, then the total length of the code is A + 1 + c and its value is mΧA + R - (2c - m), where R is the c-bit integer consisting of R and the bit that follows R. An example is the code 0001xxx, for m = 14. There are no leading 1’s, so A is 0. After removing the leading zero, the c - 1 = 3 bits that follow are R = 001. Since R < 2c - m = 2, we conclude that the length of the code is 0 + 1 + (4 - 1) = 4 and its value is 001. Similarly, the code 00100xxx for the same m = 14 has A = 0 and R = 0102 = 2. In this case, R . 2c - m = 2, so the length of the code is 0 + 1 + c = 5, the value of R is 01002 = 4, and the value of the code is 14 Χ 0 + 4 - 2 = 2. The JPEG-LS method for lossless image compression (recommendation ISO/IEC CD 14495) employs the Golomb code. The family of Golomb codes has a close relative, the exponential Golomb codes which are described on page 198. It is now clear that the best value for m depends on p, and it can be shown that this value is the integer closest to -1/ log2 p or, equivalently, the value that satisfies pm . 1/2. (3.12) 3.24 Golomb Code 165 It can also be shown that in the case of a sequence of run lengths, this integer is the median of the run lengths. Thus, for p = 0.5, m should be -1/ log2 0.5 = 1. For p = 0.7, m should be 2, because -1/ log2 0.7 . 1.94, and for p = 36/37, m should be 25, because -1/ log2(36/37) . 25.29. It should also be mentioned that Gallager and van Voorhis [Gallager and van Voorhis 75] have refined and extended Equation (3.12) into the more precise relation pm + pm+1 . 1 < pm + pm-1. (3.13) They proved that the Golomb code is the best prefix code when m is selected by their inequality. We first show that for a given p, inequality (3.13) has only one solution m. We manipulate this inequality in four steps as follows: pm(1 + p) . 1 < pm-1(1 + p), pm . 1 (1 + p) < pm-1, m . 1 log p log 1 1 + p >m- 1, m . - log(1 + p) log p >m- 1, from which it is clear that the unique value of m is m = $- log2(1 + p) log2 p %. (3.14) Three examples are presented here to illustrate the performance of the Golomb code in compressing run lengths. The first example is the binary string (3.11), which has 41 zeros and 18 ones. The probability of a zero is therefore 41/(41 + 18) . 0.7, yielding m = - log 1.7/ log 0.7 = 1.487 = 2. The sequence of run lengths 5, 2, 0, 3, 1, 6, 0, 0, 1, 3, 5, 3, 2, 3, 0, 1, 4, and 2 can therefore be encoded with the Golomb codes for m = 2 into the string of 18 codes 1101|100|00|101|01|11100|00|00|01|101|1101|101|100|101|00|01|1100|100. The result is a 52-bit string that compresses the original 59 bits. There is almost no compression because p isn’t large. Notice that string (3.11) has three short runs of 1’s, which can be interpreted as four empty runs of zeros. It also has three runs (of zeros) of length 1. The next example is the 94-bit string 00000000001000000000100000001000000000001000000001000000000000100000000100000001000000000010000000, which is sparser and therefore compresses better. It consists of 85 zeros and 9 ones, so p = 85/(85 + 9) = 0.9. The best value of m is therefore m = - log(1.9)/ log(0.9) = 6.09 = 7. The 10 runs of zeros have lengths 10, 9, 7, 11, 8, 12, 8, 7, 10, and 7. When encoded by the Golomb codes for m = 7, the run lengths become the 47-bit string 10100|10011|1000|10101|10010|10110|10010|1000|10100|1000, resulting in a compression factor of 94/47 = 2. 166 3. Advanced VL Codes The third, extreme, example is a really sparse binary string that consists of, say, 106 bits, of which only 100 are ones. The probability of zero is p = 106/(106+102) = 0.9999, implying m = 6932. There are 101 runs, each about 104 zeros long. The Golomb code of 104 for m = 6932 is 14 bits long, so the 101 runs can be compressed to 1414 bits, yielding the impressive compression factor of 707! In summary, given a binary string, we can employ the method of run-length encoding to compress it with Golomb codes in the following steps: (1) count the number of zeros and ones, (2) compute the probability p of a zero, (3) use Equation (3.14) to compute m, (4) construct the family of Golomb codes for m, and (5) for each run-length of n zeros, write the Golomb code of n on the compressed stream. In order for the run lengths to be meaningful, p should be large. Small values of p, such as 0.1, result in a string with more 1’s than zeros and thus in many short runs of zeros and long runs of 1’s. In such a case, it is possible to use RLE to compress the runs of 1’s. In general, we can talk about a binary string whose elements are r and s (for run and stop). For r, we should select the more common element, but it has to be very common (the distribution of r and s should be skewed) for RLE to produce good compression. Values of p around 0.5 result in runs of both zeros and 1’s, so regardless of which bit is selected for the r element, there will be many runs of length zero. For example, the string 00011100110000111101000111 has the following run lengths of zeros 3, 0, 0, 2, 0, 4, 0, 0, 0, 1, 3, 0, 0 and similar run lengths of 1’s 0, 0, 3, 0, 2, 0, 0, 0, 4, 1, 0, 0, 3. In such a case, RLE is not a good choice for compression and other methods should be considered. Another approach to adaptive RLE is to use the binary string input so far to estimate p and from it to estimate m, and then use the new value of m to encode the next run length (not the current one because the decoder cannot mimic this). Imagine that three runs of 10, 15, and 21 zeros have been input so far, and the first two have already been compressed. The current run of 21 zeros is first compressed with the current value of m, then a new p is computed as (10 + 15 + 21)/[(10 + 15 + 21) + 3] and is used to update m either from -1/ log2 p or from Equation (3.14). (The 3 is the number of 1’s input so far, one for each run.) The new m is used to compress the next run. The algorithm accumulates the lengths of the runs in variable L and the number of runs in N. Figure 3.53 is a simple pseudocode listing of this method. (A practical implementation should halve the values of L and N from time to time, to prevent them from overflowing.) In addition to the codes, Solomon W. Golomb has his “own” Golomb constant: 0.624329988543550870992936383100837244179642620180529286. 3.25 Rice Codes The Golomb codes constitute a family that depends on the choice of a parameter m. The case where m is a power of 2 (m = 2k) is special and results in a Rice code (sometimes also referred to as Golomb–Rice code), so named after its originator, Robert F. Rice ([Rice 79], [Rice 91], and [Fenwick 96a]). The Rice codes are also related to the subexponential code of Section 3.26. A Rice code depends on the choice of a base 3.25 Rice Codes 167 L = 0; % initialize N = 0; m = 1; % or ask user for m % main loop for each run of r zeros do construct Golomb code for r using current m. write it on compressed stream. L = L + r: % update L, N, and m N = N + 1; p = L/(L + N); m = -1/ log2 p + 0.5; endfor; Figure 3.53: Simple Adaptive Golomb RLE Encoding. k and is computed in the following steps: (1) Separate the sign bit from the rest of the number. This is optional and the bit becomes the most-significant bit of the Rice code. (2) Separate the k LSBs. They become the LSBs of the Rice code. (3) Code the remaining j = n/2k bits as either j zeros followed by a 1 or j 1’s followed by a 0 (similar to the unary code). This becomes the middle part of the Rice code. Thus, this code is computed with a few logical operations, which is faster than computing a Huffman code, a process that requires sliding down the Huffman tree while collecting the individual bits of the code. This feature is especially important for the decoder, which has to be simple and fast. Table 10.28 shows examples of this code for k = 2, which corresponds to m = 4 (the column labeled “No. of ones” lists the number of 1’s in the middle part of the code). Notice that the codes of this table are identical (except for the extra, optional, sign bit) to the codes on the 3rd row (the row for m = 4) of Table 3.50. No. of i Binary Sign LSB ones Code i Code 0 0 0 00 0 0|0|00 1 1 0 01 0 0| 0|01 -1 1|0|01 2 10 0 10 0 0| 0|10 -2 1|0|10 3 11 0 11 0 0| 0|11 -3 1|0|11 4 100 0 00 1 0|10|00 -4 1|10|00 5 101 0 01 1 0|10|01 -5 1|10|01 6 110 0 10 1 0|10|10 -6 1|10|10 7 111 0 11 1 0|10|11 -7 1|10|11 8 1000 0 00 2 0|110|00 -8 1|110|00 11 1011 0 11 2 0|110|11 -11 1|110|11 12 1100 0 00 3 0|1110|00 -12 1|1110|00 15 1111 0 11 3 0|1110|11 -15 1|1110|11 Table 3.54: Various Positive and Negative Rice Codes. 168 3. Advanced VL Codes The length of the (unsigned) Rice code of the integer n with parameter k is 1+k+ n/2k bits, indicating that these codes are suitable for data where the integer n appears with a probability P(n) that satisfies log2 P(n) = -(1 + k + n/2k) or P(n) . 2-n, an exponential distribution, such as the Laplace distribution. (See [Kiely 04] for a detailed analysis of the optimal values of the Rice parameter.) The Rice code is instantaneous, once the decoder reads the sign bit and skips to the first 0 from the left, it knows how to generate the left and middle parts of the code. The next k bits should be read and appended to that. Rice codes are ideal for data items with a Laplace distribution, but other prefix codes exist that are easier to construct and to decode and that may, in certain circumstances, outperform the Rice codes. Table 10.29 lists three such codes. The “pod” code, due to Robin Whittle [firstpr 06], codes the number 0 with the single bit 1, and codes the binary number 1 b...b  k as 0...0  k+1 1 b...b  k . In two cases, the pod code is one bit longer than the Rice code, in four cases it has the same length, and in all other cases it is shorter than the Rice codes. The Elias gamma code [Fenwick 96a] is identical to the pod code minus its leftmost zero. It is therefore shorter, but does not provide a code for zero (see also Table 3.10). The biased Elias gamma code corrects this fault in an obvious way, but at the cost of making some codes one bit longer. Number Pod Elias Biased Elias Dec Binary gamma gamma 0 00000 1 1 1 00001 01 1 010 2 00010 0010 010 011 3 00011 0011 011 00100 4 00100 000100 00100 00101 5 00101 000101 00101 00110 6 00110 000110 00110 00111 7 00111 000111 00111 0001000 8 01000 00001000 0001000 0001001 9 01001 00001001 0001001 0001010 10 01010 00001010 0001010 0001011 11 01011 00001011 0001011 0001100 12 01100 00001100 0001100 0001101 13 01101 00001101 0001101 0001110 14 01110 00001110 0001110 0001111 15 01111 00001111 0001111 000010000 16 10000 0000010000 000010000 000010001 17 10001 0000010001 000010001 000010010 18 10010 0000010010 000010010 000010011 Table 3.55: Pod, Elias Gamma, and Biased Elias Gamma Codes. There remains the question of what base value n to select for the Rice codes. The base determines how many low-order bits of a data symbol are included directly in the 3.25 Rice Codes 169 Rice code, and this is linearly related to the variance of the data symbol. Tony Robinson, the developer of the Shorten method for audio compression [Robinson 94], provides the formula n = log2[log(2)E(|x|)], where E(|x|) is the expected value of the data symbols. This value is the sum |x|p(x) taken over all possible symbols x. Figure 3.56 lists the lengths of various Rice codes and compares them to the length of the standard binary (beta) code. 10 100 1000 10000 10 20 40 60 80 100 120 5 106 k=2 n 3 4 5 k=8 k=12 k=16 Binary Code Length (* Lengths of binary code and 7 Rice codes *) bin[i_] := 1 + Floor[Log[2, i]]; Table[{Log[10, n], bin[n]}, {n, 1, 1000000, 500}]; gb = ListPlot[%, AxesOrigin -> {0, 0}, PlotJoined -> True, PlotStyle -> { AbsoluteDashing[{6, 2}]}] rice[k_, n_] := 1 + k + Floor[n/2^k]; k = 2; Table[{Log[10, n], rice[k, n]}, {n, 1, 10000, 10}]; g2 = ListPlot[%, AxesOrigin -> {0, 0}, PlotJoined -> True] k = 3; Table[{Log[10, n], rice[k, n]}, {n, 1, 10000, 10}]; g3 = ListPlot[%, AxesOrigin -> {0, 0}, PlotJoined -> True] k = 4; Table[{Log[10, n], rice[k, n]}, {n, 1, 10000, 10}]; g4 = ListPlot[%, AxesOrigin -> {0, 0}, PlotJoined -> True] k = 5; Table[{Log[10, n], rice[k, n]}, {n, 1, 10000, 10}]; g5 = ListPlot[%, AxesOrigin -> {0, 0}, PlotJoined -> True] k = 8; Table[{Log[10, n], rice[k, n]}, {n, 1, 100000, 50}]; g8 = ListPlot[%, AxesOrigin -> {0, 0}, PlotJoined -> True] k = 12; Table[{Log[10, n], rice[k, n]}, {n, 1, 500000, 100}]; g12 = ListPlot[%, AxesOrigin -> {0, 0}, PlotJoined -> True] k = 16; Table[{Log[10, n], rice[k, n]}, {n, 1, 1000000, 100}]; g16 = ListPlot[%, AxesOrigin -> {0, 0}, PlotJoined -> True] Show[gb, g2, g3, g4, g5, g8, g12, g16, PlotRange -> {{0, 6}, {0, 120}}] Figure 3.56: Lengths of Various Rice Codes. 170 3. Advanced VL Codes 3.26 Subexponential Code The subexponential code of this section is related to the Rice codes. Like the Golomb codes and the Rice codes, the subexponential code depends on a parameter k . 0. The main feature of the subexponential code is its length. For integers n < 2k+1, the code length increases linearly with n, but for larger values of n, it increases logarithmically. The subexponential code of the nonnegative integer n is computed in two steps. In the first step, values b and u are calculated by b = k, if n < 2k, log2 n, if n . 2k, and u = 0, if n < 2k, b - k + 1, if n . 2k. In the second step, the unary code of u (in u+1 bits) is followed by the b least-significant bits of n to become the subexponential code of n. Thus, the total length of the code is u+1+b = k + 1, if n < 2k, 2log2 n - k + 2, if n . 2k. Table 7.139 lists examples of the subexponential code for various values of n and k. It can be shown that for a given n, the code lengths for consecutive values of k differ by at most 1. n k= 0 k = 1 k = 2 k = 3 k = 4 k = 5 0 0| 0|0 0|00 0|000 0|0000 0|00000 1 10| 0|1 0|01 0|001 0|0001 0|00001 2 110|0 10| 0 0|10 0|010 0|0010 0|00010 3 110|1 10| 1 0|11 0|011 0|0011 0|00011 4 1110|00 110|00 10|00 0|100 0|0100 0|00100 5 1110|01 110|01 10|01 0|101 0|0101 0|00101 6 1110|10 110|10 10|10 0|110 0|0110 0|00110 7 1110|11 110|11 10|11 0|111 0|0111 0|00111 8 11110|000 1110|000 110|000 10|000 0|1000 0|01000 9 11110|001 1110|001 110|001 10|001 0|1001 0|01001 10 11110|010 1110|010 110|010 10|010 0|1010 0|01010 11 11110|011 1110|011 110|011 10|011 0|1011 0|01011 12 11110|100 1110|100 110|100 10|100 0|1100 0|01100 13 11110|101 1110|101 110|101 10|101 0|1101 0|01101 14 11110|110 1110|110 110|110 10|110 0|1110 0|01110 15 11110|111 1110|111 110|111 10|111 0|1111 0|01111 16 111110|0000 11110|0000 1110|0000 110|0000 10|0000 0|10000 Table 3.57: Some Subexponential Codes. Subexponential codes are used in the progressive FELICS method for image compression [Howard and Vitter 94a]. 3.27 Codes Ending with “1” 171 3.27 Codes Ending with “1” In general, the particular bits that constitute a code are irrelevant. Given a set of codewords that have the desired properties, we don’t check the individual bits of a code and object if there is a dearth of zeros or too few 1’s. Similarly, we don’t complain if most or all of the codes start with a 1 or end with a 0. The important requirements of variable-length codes are (1) to have a set of codes that feature the shortest average length for a given statistical distribution of the source symbols and (2) to have a UD code. However, there may be applications where it is advantageous to have codes that start or end in a special way, and this section presents prefix codes that end with a 1. The main contributors to this line of research are R. Capocelli, A. De Santis, T. Berger, and R. Yeung (see [Capocelli and De Santis 94] and [Berger and Yeung 90]). They discuss special applications and coin the term “feasible codes” for this type of prefix codes. They also show several ways to construct such codes and prove the following bounds on their average lengths. The average length E of an optimum feasible code for a discrete source with entropy H satisfies H + pN . E . H + 1.5, where pN is the smallest probability of a symbol from the source. This section discusses a simple way to construct a set of feasible codes from a set of Huffman codes. The main idea is to start with a set of Huffman codes and append a 1 to each code that ends with a 0. Such codes are called “derived codes.” They are easy to construct, but are not always the best feasible codes (they are not optimal). The construction of the codes is as follows: Given a set of symbols and their probabilities, construct their Huffman codes. The subset of codes ending with a 0 is denoted by C0 and the subset of codes ending with a 1 is denoted by C1. We also denote by p(C0) and p(C1) the sum of probabilities of the symbols in the two subsets, respectively. Notice that p(C0)+p(C1) = 1, which implies that either p(C0) or p(C1) is less than or equal to 1/2 and the other one is greater than or equal to 1/2. If p(C0) . 1/2, the derived code is constructed by appending a 1 to each codeword in C0; the codewords in C1 are not modified. If, on the other hand, p(C0) > 1/2, then the zeros and 1’s are interchanged in the entire set of Huffman codes, resulting in a new p(C0) that is less than or equal to 1/2, and the derived code is constructed as before. As an example, consider the set of six symbols with probabilities 0.26, 0.24, 0.14, 0.13, 0.12, and 0.11. The entropy of this set is easily computed at approximately 2.497 and one set of Huffman codes for these symbols (from high to low probabilities) is 11, 01, 101, 100, 001, and 000. The average length of this set is 0.26Χ2 + 0.24Χ2 + 0.14Χ3 + 0.13Χ3 + 0.12Χ3 + 0.11Χ3 = 2.5 bits. It is easy to verify that p(C0) = 0.13 + 0.11 = 0.24 < 0.5, so the derived code becomes 11, 01, 101, 1001, 001, and 0001, with an average size of 2.74 bits. On the other hand, if we consider the set of Huffman codes 00, 10, 010, 011, 110, and 111 for the same symbols, we compute p(C0) = 0.26 + 0.24 + 0.14 + 0.12 = 0.76 > 0.5, so we have to interchange the bits and append a 1 to the codes of 0.13 and 0.11. This again results in a derived code with an average length E of 2.74, which satisfies E 2, is nonsynchronous. Also, for any n . 12, there is a source where all the Huffman codes are of this type. The greatest common divisor (gcd) of a set of nonzero integers li is the largest integer that divides each of the li evenly (without remainders). If the gcd of a set is 1, then the members of the set are relatively prime. An example of relatively prime integers is (9, 28, 29). These criteria suggest that many Huffman codes are nonsynchronous and cannot even be modified to become synchronous. However, there are also many Huffman codes that are either synchronous or can be twisted to become synchronous. We start with the concept of quantized source. Given a source S of data symbols with known probabilities, we construct a Huffman code for the source. We denote the length of the longest codeword by L and denote by .i the number of codewords with length i. The vector S = (.1, .2, . . . , .L) is called the quantized representation of the source S. A source is said to be gapless if the first nonzero .i is followed only by nonzero .j (i.e., if all the zero .’s are concentrated at the start of the quantized vector). The following are criteria for synchronous Huffman codes: 1. If .1 is positive (i.e., there are some codewords with length 1), then any Huffman code for S is synchronous. 2. Given a gapless source S where the minimum codeword length is 2, then S has a synchronous Huffman code, unless its quantized representation is one of the following (0, 4), (0, 1, 6), (0, 1, 1, 10), and (0, 1, 1, 1, 18). 4.3 Synchronous Prefix Codes 189 The remainder of this section is based on [Capocelli and De Santis 92], a paper that offers an advanced discussion of synchronous prefix codes, with theorems and proofs. We present only a short summary of the many results discovered by these authors. Given a feasible code (Section 3.27) where the length of the longest codeword is L, all the codewords of the form 00 . . . 0    L-1 1, 00 . . . 0    L-2 1, and 1 00 . . . 0    L-2 1, are synchronizing codewords. If there are no such codewords, it is possible to modify the code as follows. Select a codeword of length L with the most number of consecutive zeros on the left 00 . . . 0    j 1x . . . x and change the 1 to a 0 to obtain 00 . . . 0    j+1 x . . . x. Continue in this way, until all the x bits except the last one have been changed to 1’s and the codeword has the format 00 . . . 0    L-1 1. This codeword is now synchronizing. As an example, given the feasible code 0001, 001, 0101, 011, 1001, 101, and 11, L = 4 and the three codewords in boldface are synchronizing. This property can be extended to nonbinary codes, with the difference that instead of a 1, any digit other than 0 can be substituted. Thus, given the feasible ternary code 00001, 0001, 0002, 001, 002, 01, 02, 1001, 101, 102, 201, 21, and 22, L equals 5 and the first three codewords are synchronizing. Going back to the definition of a synchronizing codeword, it is easy to show by direct checks that the first 11 codewords of this code satisfy this definition and are therefore synchronizing. Table 4.9 (after [Capocelli and De Santis 92]) lists a 4-ary code for the 21-character Latin alphabet. The rules above imply that all the codewords that end with 3 are synchronizing, while our previous definition shows that 101 and 102 are also synchronizing. Letter Freq. Code Letter Freq. Code Letter Freq. Code h 5 0000 d 17 *101 r 67 20 x 6 0001 l 21 *102 s 68 21 v 7 0002 p 30 *103 a 72 22 f 9 100 c 33 11 t 72 *03 b 12 001 m 34 12 u 74 *13 q 13 002 o 44 01 e 92 *23 g 14 *003 n 60 02 i 101 *3 Table 4.9: Optimal 4-ary Synchronous Code for the Latin Alphabet. See also the code of Section 3.21.1. 190 4. Robust VL Codes 4.4 Resynchronizing Huffman Codes Because of the importance and popularity of Huffman codes, much research has gone into every aspect of these codes. The work described here, due to [Rudner 71], shows how to identify a resynchronizing Huffman code (RHC), a set of codewords that allows the decoder to always synchronize itself following an error. Such a set contains at least one synchronizing codeword (SC) with the following property: any bit string followed by an SC is a sequence of valid codewords. 0111 0110 010 0011 0010 111 110 000 10 0 0 0 0 0 0 1 1 1 1 1 1 Figure 4.10: A Resynchronizing Huffman Code Tree. Figure 4.10 is an example of such a code. It is a Huffman code tree for nine symbols having probabilities of occurrence of 1/4, 1/8, 1/8, 1/8, 1/8, 1/16, 1/16, 1/16, and 1/16. Codeword 010 (underlined) is an SC. A direct check verifies that any bit string followed by 010 is a sequence of valid codewords from this tree. For example, 001010001|010 is the string of four codewords 0010|10|0010|10. If an encoded file has an error at point A and an SC appears later, at point B, then the bit string from A to B (including the SC) is a string of valid codewords (although not the correct codewords) and the decoder becomes synchronized at B (although it may miss decoding the SC itself). Reference [Rudner 71] proposes an algorithm to construct an RHC, but the lengths of the codewords must satisfy certain conditions. Given a set of prefix codewords, we denote the length of the shortest code by m. An integer q < m is first determined by a complex test (listed below) that goes over all the nonterminal nodes in the code tree and counts the number of codewords of the same length that are descendants of the node. Once q has been computed, the SC is the codeword 0q1m-q0q. It is easy to see that m = 2 for the code tree of Figure 4.10. The test results in q = 1, so the SC is 010. In order for an SC to exist, the code must either have m = 1 (in which case the SC has the form 02q) or must satisfy the following conditions: 1. The value of m is 2, 3, or 4. 2. The code must contain codewords with all the lengths from m to 2m- 1. 3. The value of q that results from the test must be less than m. Figure 4.11 is a bigger example with 20 codewords. The shortest codeword is three bits long, so m = 3. The value of q turns out to be 2, so the SC is 00100. A random bit string followed by the SC, such as 01000101100011100101010100010101|00100, becomes the string of nine codewords 0100|01011|00011|100|1010|1010|00101|0100|100. The test to determine the value of q depends on many quantities that are defined in [Rudner 71] and are too specialized to be described here. Instead, we copy this test verbatim from the reference. 4.4 Resynchronizing Huffman Codes 191 1 1 2 3 4 5 6 1 1 1 1 1 1 1 1 1 1 1 0 0 0 0 0 0 00011 00011 00100 00100 00101 00101 00110 00110 00111 00111 01010 01011 01010 01011 0 000100 000100 000101 000101 0 0 0 0 1111 1111 1110 1110 1101 1101 1100 1100 1011 1011 1010 1010 0111 0111 0110 0110 0100 0100 100 100 0000 0000 0 0 Figure 4.11: A Resynchronizing Huffman Code Tree With 20 Codewords. “Let index I have 2 . m . 4 and p < m. Let Qj(h) = (h +m- j) + j x=j-m+2 cx-h-1,x where ca,b = 0 for a . 0. Let q be the smallest integer p . q < m, if any exists, for which nj . Qj(q) for all j, m . j . M. If no such integer exists, let q = max(m, p). Then r0 . q +m.” The operation of an SC is especially simple when it follows a nonterminal node in the code tree. Checking Figure 4.11 verifies that when any interior node is followed by the SC, the result is either one or two codewords. In the former case, the SC is simply a suffix of the single resulting codeword, while in the latter case, some of the leftmost zeros of the SC complete the string of the interior node, and the remainder of the SC (a string of the form 0k1m-q0q, where k is between 0 and q - 1) is a valid codeword that is referred to as a reset word. In the code tree of Figure 4.11, the reset words are 0100 and 100. When the interior node 0101 is followed by the SC, the result is 01010|0100. When 0 is followed by the SC, the result is 000100, and when 01 is followed by 00100, we obtain 0100|100. Other codewords may also have this property (i.e., they may reset certain interior nodes). This happens when a codeword is a suffix of another codeword, such as 0100, which is a suffix of 000100 in the tree of Figure 4.11. Consider the set that consists of the SC (whose length is m + q bits) and all the reset words (which are shorter). This set can reset all the nonterminal nodes, which implies a certain degree of nonuniformity in the code tree. The tree of Figure 4.11 has levels from 1 to 6, and a codeword at level D is D bits long. Any node (interior or not) at level D must therefore have at least one codeword of length L that satisfies either D + 1 . L . D + q - 1 (this corresponds to the case where appending the SC to the node results in two codewords) or L = D + q + m (this corresponds to the case where appending the SC to the node results in one codeword). Thus, the lengths of codewords in the tree are restricted and may not truely reflect the data symbols’ probabilities. 192 4. Robust VL Codes Another aspect of this restriction is that the number of codewords of the same length that emanate from a nonterminal node is at most 2q. Yet another nonuniformity is that short codewords tend to concentrate on the 0-branch of the tree. This downside of the RHC has to be weighed against the advantage of having a resynchronizing code. Another important factor affecting the performance of an RHC is the expected time to resynchronize. From the way the Huffman algorithm works, we know that a codeword of length L appears in the encoded stream with probability 1/2L. The length of the SC is m+q, so if we consider only the resynchronization provided by the SC, and ignore the reset words, then the expected time . to resynchronize (the inverse of the probability of occurrence of the SC) is 2m+q. This is only an upper bound on . , because the reset words also contribute to synchronization and reduce the expected time . . If we assume that the SC and all the reset words reset all the nonterminal (interior) nodes, then it can be shown that 1 2m-1 - 2m+q is a lower bound on . . Better estimates of . can be computed for any given code tree by going over all possible errors, determining the recovery time .i for each possible error i, and computing the average . . An RHC allows the decoder to resynchronize itself as quickly as possible following an error, but there still are two points to consider: (1) the number of data symbols decoded while the decoder is not synchronized may differ from the original number of symbols and (2) the decoder does not know that a slippage occurred. In certain applications, such as run-length encoding of bi-level images, where runs of identical pixels are decoded and placed consecutively in the decoded image, these points may cause a noticeable error when the decoded image is later examined. A possible solution, due to [Lam and Kulkarni 96], is to take an RHC and modify it by including an extended synchronizing codeword (ESC), which is a bit pattern that cannot appear in any concatenation of codewords. The resulting code is no longer an RHC, but is still an RVLC (reversible VLC, page 195). The idea is to place the ESC at regular intervals in the encoded bitstream, say, after every 10 symbols or at the end of each line of text or each row of pixels. When the decoder decodes an ESC, it knows that 10 symbols should have been decoded since the previous ESC. If this is not true, the decoder can issue an error message, warning the user of a potential synchronization problem. In order to understand the particular construction of an ESC, we consider all the cases in which a codeword c can be decoded incorrectly when there are errors preceding it. Luckily, there are only two such cases, illustrated in Figure 4.12. (a) (b) c c Figure 4.12: Two Ways to Incorrectly Decode a Codeword. 4.5 Bidirectional Codes 193 In part (a) of the figure, a prefix of c is decoded as the suffix of some other codeword and a suffix of c is decoded as the prefix of another codeword (in between, parts of c may be decoded as several complete codewords). In part (b), codeword c is concatenated with some bits preceding it and some bits following it, to form a valid codeword. This simple analysis points the way toward an ESC, because it is now clear that this bit pattern should satisfy the following conditions: 1. If the ESC can be written as the concatenation of two bit strings .ί where . is a suffix of some codeword, then ί should not be the prefix of any other codeword. Furthermore, if ί itself can be written as the concatenation of two bit strings .., then . can be empty or a concatenation of some codewords, and . should be nonempty, should not be the prefix of any codeword, and no codeword should be a prefix of .. This takes care of case (a). 2. The ESC should not be a substring of any codeword. This takes care of case (b). If the ESC satisfies these conditions, then the decoder can recognize it even in the presence of errors preceding it. Once the decoder recognizes the ESC, it reads the bits that follow until it can continue the decoding. As can be expected, the life of a researcher in the field of coding is never that simple. It may happen that an error corrupts part of the encoded bit stream into an ESC. Also, an error can corrupt the ESC pattern itself. In these cases, the decoder loses synchronization, but regains it when the next ESC is found later. It is also possible to append a fixed-length count to each ESC, such that the first ESC is followed by a count of 1, the second ESC is followed by a 2, and so on. If the latest ESC had a count i - 1 and the next ESC has a count of i + 1, then the decoder knows that the ESC with a count of i has been missed because of such an error. If the previous ESC had a count of i and the current ESC has a completely different count, then the current ESC is likely spurious. Given a Huffman code tree, how can an ESC be constructed? We first observe that it is always possible to rearrange a Huffman code tree such that the longest codeword (the codeword of the least-probable symbol) consists of all 1’s. If this codeword is 1k, then we construct the ESC as the bit pattern 1k1k-10 = 12k-10. This pattern is longer than the longest codeword, so it cannot be a substring of any codeword, thereby satisfying condition 2 above. For condition 1, consider the following. All codeword suffixes that consist of consecutive 1’s can have anywhere from no 1’s to (k-1) 1’s. Therefore, if any prefix of the ESC is the suffix of a codeword, the remainder of the ESC will be a string of the form 1j0 where j is between k and 2k - 1, and such a string is not the prefix of any codeword, thereby satisfying condition 1. 4.5 Bidirectional Codes This book is enlivened by the use of quotations and epigraphs. Today, it is easy to search the Internet and locate quotations on virtually any topic. Many web sites maintain large collections of quotations, there are many books whose full text can easily be searched, and the fast, sophisticated Internet search engines can locate many occurrences of any word or phrase. Before the age of computers, the task of finding words, phrases, and full quotations required long visits to a library and methodical reading of many texts. For this reason, 194 4. Robust VL Codes scholars sometimes devoted substantial parts of their working careers to the preparation of a concordance. A concordance is an alphabetical list of the principal words found in a book or a body of work, with their locations (page and line numbers) and immediate contexts. Common words such as “of” and “it” are excluded. The task of manually compiling and publishing a concordance was often gargantuan, which is why concordances were generated only for highly-valued works, such as the Bible, the writings of St. Thomas Aquinas, the Icelandic sagas, and the works of Shakespeare and Wordsworth. The emergence of the digital computer, in the late 1940s and early 1950s, has given a boost to the art of concordance compiling. Suddenly, it became possible to store an entire body of literature in the computer and point to all the important words in it. It was also in the 1950s that Hans Peter Luhn, a computer scientist at IBM, conceived the technique of KWIC indexing. The idea was to search a concordance for a given word and display or print a condensed list of all the occurrences of the word, each with its location (page and line numbers) and immediate context. KWIC indexing remained popular for many years until it became superseded by the full text search that so many of us take for granted. KWIC is an acronym for Key Word In Context, but like so many acronyms it has other meanings such as Kitchen Waste Into Compost, Kawartha World Issues Center, and Kids’ Well-being Indicators Clearinghouse. The following KWIC example lists the first 15 results of searching for the word may in a large collection (98,000 items) of excerpts from academic textbooks and introductory books located at the free concordance [amu 06]. No: Line: Concordance 1: 73: in the National Grid - which may accompany the beginning or end of 2: 98: of a country, however much we may analyze it into separate rules, 3: 98: and however much the analysis may be necessary to our understanding 4: 104: y to interpret an Act a court may sometimes alter considerably the 5: 110: ectly acquainted with English may know the words of the statute, but 6: 114: ime. Therefore in a sense one may speak of the Common Law as unwritten 7: 114: e for all similar cases which may arise in the future. This binding 8: 120: court will not disregard. It may happen that a question has never 9: 120: ven a higher court, though it may think a decision of a lower court 10: 122: ween the parties. The dispute may be largely a question of fact. 11: 122: nts must be delivered, and it may often be a matter of doubt how far 12: 138: of facts. The principles, it may be, give no explicit answer to the 13: 138: the conclusions of a science may be involved in its premisses, and 14: 144: at conception is that a thing may change and yet remain the same thing. 15: 152: ions is unmanageable, Statute may undertake the work of codification, Thus, a computerized concordance that supports KWIC searches consists of a set of texts and a dictionary. The dictionary lists all the important words in the texts, each with pointers to all its occurrences. Because of their size, the texts are normally stored in compressed format where each character is assigned a variable-length code (perhaps a Huffman code). The software inputs a search term, such as may, from the user, locates it in the dictionary, follows the first pointer to the text, and reads the immediate context of the first occurrence of may. The point is that this context both precedes and follows the word, but Huffman codes (and variable-length codes in general) are designed to be read 4.5 Bidirectional Codes 195 and decoded from left to right (from early to late). In general, when reading a group of variable-length codes, it is impossible to reverse direction and read the preceding code, because there is no way to tell its length. We say that variable-length codes are unidirectional, but there are applications where bidirectional (or reversible) codes are needed. Computerized concordances and KWIC indexing is one such application. Another application, no less important, is data integrity. We already know, from the discussion of synchronous codes in Section 4.3, that errors may occur and they tend to propagate through a sequence of variable-length codes. One way to limit error propagation (slippage) is to organize a compressed file in records, where each record consists of several variable-length codes. The record is read and decoded from beginning to end, but if an error is discovered, the decoder tries to read the record from end to start. If there is only one error in the record, the two readings and decodings can sometimes be combined to isolate the error and perhaps even to correct it There are other, much less important applications of bidirectional codes. In the early days of the digital computer, magnetic tapes were the main input/output devices. A tape is sequential storage that lends itself to linear reading and writing. If we want to read an early record from a tape, we generally have to rewind the tape and then skip forward to the desired record. The ability to read a tape backward can speed up tape input/output and make the tape look more like a random-access I/O device. Another example of the use of bidirectional codes is a little-known data structure called deque (short for double-ended queue, and sometimes spelled “dequeue”). This is a linear data structure where elements can be added to or deleted from either end. (This is in contrast to a queue, where elements can only be added to the head and deleted from the tail.) If the elements of the deque are compressed by means of variable-length codes, then bidirectional codes allow for easy access of the structure from either end. A good programmer is someone who always looks both ways before crossing a one-way street. —Doug Linder It should be noted that fixed-length codes are bidirectional, but they generally do not provide any compression (the Tunstall code of Section 2.6 is an exception). We say that fixed-length codes are trivially bidirectional. Thus, we need to develop variablelength codes that are as short as possible on average but are also bidirectional. The latter requirement is important, because the bidirectional codes are going to be used for compression (where they are often called reversible VLCs or RVLCs). Anyone with even little experience in the field of variable-length codes will immediately realize that it is easy to design bidirectional codes. Simply dedicate a bit pattern p to be both the prefix and suffix of all the codes, and make sure that p does not appear inside a code. Thus, if p = 101, then 101|10011001001|101 is an example of such a code. It is easy to see that such codes can be read in either direction, but it is also obvious that they are too long, because the code bits between the suffix and prefix are restricted to patterns that do not contain p. A little thinking shows that p doesn’t even have to appear twice in each code. A code where each codeword ends with p is bidirectional. A string of such codes has the form papbpc . . . pypz and it is obvious that it can be read in either direction. The taboo codes of Section 3.16 are an example of this type of variable-length code. 196 4. Robust VL Codes The average length of a code is therefore important even in the case of bidirectional codes. Recall that the most common variable-length codes can be decoded easily and uniquely because they are prefix codes and therefore instantaneous. This suggests the idea of constructing a suffix code, a code where no codeword is the suffix of another codeword. A suffix code can be read in reverse and decoded uniquely in much the same way that a prefix code is uniquely decoded. Thus, a code that is both a prefix code and a suffix code is bidirectional. Such a code is termed affix (although some authors use the terms “biprefix” and “never-self-synchronizing”). The next few paragraphs are based on [Fraenkel and Klein 90], a work that analyses Huffman codes, looking for ways to construct affix Huffman codes or modify existing Huffman codes to make them affix and thus bidirectional. The authors first show how to start with a given affix Huffman code C and double its size. The idea is to take every codeword ci in C and create two new codewords from it by appending a bit to it. The two new codewords are therefore ci0 and ci1. The resulting set of codewords is affix as can be seen from the following argument. 1. No other codeword in C starts with ci (ci is not the prefix of any other codeword). Therefore, no other codeword in the new code starts with ci. As a result, no other codeword in the new code starts with ci0 or ci1. 2. Similarly, ci is not the suffix of any other codeword in C, therefore neither ci0 nor ci1 are suffixes of a codeword in the new code. Given the affix Huffman code 01, 000, 100, 110, 111, 0010, 0011, 1010, and 1011, we apply this method to double it and construct the code 010, 0000, 1000, 1100, 1110, 00100, 00110, 10100, 011, 0001, 1001, 1101, 1111, 00101, 00111, and 10101. A simple check verifies that the new code is affix. The conclusion is that there are infinitely many affix Huffman codes. On the other hand, there are cases where affix Huffman codes do not exist. Consider, for example, codewords of length 1. In the trivial case where there are only two codewords, each is a single bit. This case is trivial and is also a fixed-length code. Thus, a variable-length code can have at most one codeword of length 1 (a single bit). If a code has such a codeword, then it is the suffix of other codewords, because in a complete prefix code, codewords must end with both 0 and 1. Thus, a code one of whose codewords is of length 1 cannot be affix. Such Huffman codes exist for sets of symbols with skewed probabilities. In fact, it is known that the existence of a codeword ci of length 1 in a Huffman code implies that the probability of the symbol that corresponds to ci must be greater than 1/3. The authors then describe a complex algorithm (not listed here) to construct affix Huffman codes for cases where such codes exist. There are many other ways to construct RVLCs from Huffman codes. Table 4.13 (after [Lakovi΄c and Villasenor 03], see also Table 4.22) lists a set of Huffman codes for the 26 letters of the English alphabet together with three RVLC codes for the same symbols. These codes were constructed by algorithms proposed by [Takishima et al. 95], [Tsai and Wu 01a], and [Lakovi΄c and Villasenor 03]. The remainder of this section describes the extension of Rice codes and exponential Golomb (EG) codes to bidirectional codes (RVLCs). The resulting bidirectional codes have the same average length as the original, unidirectional Rice and EG codes. They have been adopted by the International Telecommunications Union (ITU) for use 4.5 Bidirectional Codes 197 p Huffman Takishima Tsai Lakovi΄c E 0.14878 001 001 000 000 T 0.09351 110 110 111 001 A 0.08833 0000 0000 0101 0100 O 0.07245 0100 0100 1010 0101 R 0.06872 0110 1000 0010 0110 N 0.06498 1000 1010 1101 1010 H 0.05831 1010 0101 0100 1011 I 0.05644 1110 11100 1011 1100 S 0.05537 0101 01100 0110 1101 D 0.04376 00010 00010 11001 01110 L 0.04124 10110 10010 10011 01111 U 0.02762 10010 01111 01110 10010 P 0.02575 11110 10111 10001 10011 F 0.02455 01111 11111 001100 11110 M 0.02361 10111 111101 011110 11111 C 0.02081 11111 101101 100001 100010 W 0.01868 000111 000111 1001001 100011 G 0.01521 011100 011101 0011100 1000010 Y 0.01521 100110 100111 1100011 1000011 B 0.01267 011101 1001101 0111110 1110111 V 0.01160 100111 01110011 1000001 10000010 K 0.00867 0001100 00011011 00111100 10000011 X 0.00146 00011011 000110011 11000011 11100111 J 0.00080 000110101 0001101011 100101001 100000010 Q 0.00080 0001101001 00011010011 0011101001 1000000010 Z 0.00053 0001101000 000110100011 1001011100 1000000111 Avg. length 4.15572 4.36068 4.30678 4.25145 Table 4.13: Huffman and Three RVLC Codes for the English Alphabet. in the video coding parts of MPEG-4, and especially in the H.263v2 (also known as H.263+ or H.263 1998) and H263v3 (also known as H.263++ or H.263 2000) video compression standards [T-REC-h 06]. The material presented here is based on [Wen and Villasenor 98]. The Rice codes (Section 3.25) are a special case of the more general Golomb code, where the parameter m is a power of 2 (m = 2k). Once the base k has been chosen, the Rice code of the unsigned integer n is constructed in two steps: (1) Separate the k least-significant bits (LSBs) of n. They become the LSBs of the Rice code. (2) Code the remaining j = n/2k bits in unary as either j zeros followed by a 1 or j 1’s followed by a 0. This becomes the most-significant part of the Rice code. This code is therefore easily constructed with a few logical operations. Decoding is also simple and requires only the value of k. The decoder scans the most-significant 1’s until it reaches the first 0. This gives it the value of the most- 198 4. Robust VL Codes significant part of n. The least-significant part of n is the k bits following the first 0. This simple decoding points the way to designing a bidirectional Rice code. The second part is always k bits long, so it can be read in either direction. To also make the first (unary) part bidirectional, we change it from 111 . . . 10 to 100 . . . 01, unless it is a single bit, in which case it becomes a single 0. Table 4.14 lists several original and bidirectional Rice codes. It should be compared with Table 10.28. No. of Rice Rev. n Binary LSB 1’s Code Rice 0 0 00 0 0| 00 0|00 1 1 01 0 0| 01 0|01 2 10 10 0 0| 10 0|10 3 11 11 0 0| 11 0|11 4 100 00 1 10|00 11|00 5 101 01 1 10|01 11|01 6 110 10 1 10|10 11|10 7 111 11 1 10|11 11|11 8 1000 00 2 110|00 101|00 11 1011 11 2 110|11 101|11 12 1100 00 3 1110|00 1001|00 15 1111 11 3 1110|11 1001|11 Table 4.14: Original and Bidirectional Rice Codes. An interesting property of the Rice codes is that there are 2k codes of each length and the lengths start at k + 1 (the prefix is at least one bit, and there is a k-bit suffix). Thus, for k = 3 there are eight codes of length 4, eight codes of length 5, and so on. For certain probability distributions, we may want the number of codewords of length L to grow exponentially with L, and this feature is offered by a parametrized family of codes known as the exponential Golomb codes. These codes were first proposed in [Teuhola 78] and are also identical to the triplet (s, 1,.) of start-step-stop codes. They perform well for probability distributions that are exponential but are taller than average and have a wide tail. The exponential Golomb codes depend on the choice of a nonnegative integer parameter s (that becomes the length of the suffix of the codewords). The nonnegative integer n is encoded in the following steps: 1. Compute w = 1+n/2s. 2. Compute f2s (n) = &log2 1 + n 2s '. This is the number of bits following the leftmost 1 in the binary representation of w. 3. Construct the codeword EG(n) as the unary representation of f2s (n), followed by the f2s (n) least-significant bits in the binary representation of w, followed by the s least-significant bits in the binary representation of n. Thus, the length of this codeword is the sum l(n) = 1+2f2s (n) + s = 1+2"log2 1 + n 2s #+ s = P + s, 4.5 Bidirectional Codes 199 where P is the prefix (whose length is always odd) and s is the suffix of a codeword. Because the logarithm is truncated, the length increases by 2 each time the logarithm increases by 1, i.e., each time 1 + n/2s is a power of 2. (As a side note, it should be mentioned that the exponential Golomb codes can be further generalized by substituting an arbitrary positive integer parameter m for the expression 2s. Such codes can be called generalized exponential Golomb codes.) As an example, we select s = 1 and determine the exponential Golomb code of n = 1110 = 10112. We compute w = 1+11/2s = 6 = 1102, f2(11) = log2(11/2) = 2, and construct the three parts 110|10|1. Table 4.15 lists (in column 2) several examples of the exponential Golomb codes for s = 1. Each code has an s-bit suffix and the prefixes get longer with n. The table also illustrates (in column 3) how these codes can be modified to become bidirectional. The idea is to have the prefix start and end with 1’s, to fill the odd-numbered bit positions of the prefix (except the two extreme ones) with zeros, and to fill the even-numbered positions with bit patterns that represent increasing integers. Thus, for s = 1, the 16 (7+1)-bit codewords (i.e., the codewords for n = 14 through n = 29) have a 7-bit prefix of the form 1x0y0z1 where the bits xyz take the eight values 000 through 111. Pairs of consecutive codewords have the same prefix and differ in their 1-bit suffixes. Exp. Rev. n Golomb exp. 0 0|0 0|0 1 0|1 0|1 2 100|0 101|0 3 100|1 101|1 4 101|0 111|0 5 101|1 111|1 6 11000|0 10001|0 7 11000|1 10001|1 8 11001|0 10011|0 9 11001|1 10011|1 10 11010|0 11001|0 11 11010|1 11001|1 Table 4.15: Original and Bidirectional Exponential Golomb Codes. Here is how the decoder can read such codewords in reverse and identify them. The decoder knows the value of s, so it first reads the s-bit suffix. If the next bit (the rightmost bit of the prefix) is 0, then the entire prefix is this single bit and the suffix determines the value of n (between 0 and 2s-1). Otherwise, the decoder reads bits from right to left, isolating the bits with even indexes (shown in boldface in Table 4.15) and concatenating them from right to left, until an odd-indexed bit of 1 is found. The total number of bits read is the length P of the prefix. All the prefixes of length P differ only in their even-numbered bits, and a P-bit prefix (where P is always odd) has (P - 1)/2 such bits. Thus, there are 2(P-1)/2 groups of prefixes, each with 2s identical prefixes. The decoder identifies the particular n in a group by the s-bit suffix. 200 4. Robust VL Codes The bidirectional exponential Golomb codes therefore differ from the (original) exponential Golomb codes in their construction, but the two types have the same lengths. The magical exclusive-OR. The methods presented earlier are based on sets of Huffman, Rice, and exponential Golomb codes that are modified and restricted in order to become bidirectional. In contrast, the next few paragraphs present a method, due to [Girod 99], where any set of prefix codes Bi can be transformed to a bitstring C that can be decoded in either direction. The method is simple and requires only logical operations and string reversals. It is based on a well-known “magical” property of the exclusive-OR (XOR) logical operation, and its only downside is the addition of a few extra zero bits. First, a few words about the XOR operation and what makes it special. The logical OR operation is familiar to many. It receives two bits a and b as its inputs and it outputs one bit. The output is 1 if a or b or both are 1’s. The XOR operation is similar but it excludes the case where both inputs are 1’s. Thus, the result of (1 XOR 1) is 0. Both logical operations are summarized in the table. a 0011 b 0101 a OR b 0111 a XOR b 0110 The table also shows that the XOR of any bit a with 0 is a. What makes the XOR so useful is the following property: if c = a XOR b, then b = a XOR c and a = b XOR c. This property is easily verified and it has made the XOR very popular in many applications (see page 357 of [Salomon 03] for an interesting example). This useful property of the XOR is now exploited as follows. The first idea is to start with a string of n data symbols and encode them with a prefix code to produce a string B of n codewords B1B2 . . . Bn. Now reverse each codeword Bi to become Bi and construct the string B = B1 B2 . . . Bn . Next, compute the final result C = B XOR B. The hope is that the useful property of the XOR will enable us to decode B from C with the relation B = C XOR B. This simple scheme does not work because the relation requires knowledge of B. We don’t know string B, but we know that its components are closely related to those of string B. Thus, the following trick becomes the key to this elegant method. Denote by L the length of the longest codeword Bi, append L zeros to B and prepend L zeros to B (more than L zeros can be used, but not fewer). Now perform the operation C = (B 00 . . . 0    L ) XOR (00 . . . 0    L B). It is clear that the first L bits of C are identical to the first L bits of B. Because of the choice of L, those L bits constitute at least the first codeword B1 (there may be more than one codeword and there may also be a remainder), so we can immediately reverse it to obtain B1 . Now we can read more bits from C and XOR them with B1 to obtain more bits with parts of codewords from B. This unusual decoding procedure is best illustrated by an example. We start with five symbols a1 through a5 and assign them the variable-length prefix codes 0, 10, 111, 1101, and 1100. The string of symbols a2a1a3a5a4 is compressed by this code to the string B = 10| 0|111|1100|1101. Reversing each codeword produces B = 01|0|111|0011|1011. The longest codeword is four bits long, so we select L = 4. 4.5 Bidirectional Codes 201 We append four zeros to B and prepend four zeros to B. Exclusive-ORing the strings yields B = 1001 1111 0011 010000 B = 0000 0101 1100 111011 C = 1001 1010 1111 101011 Decoding is done in the following steps: 1. XOR the first L bits of C and B. This results in 1001 . 0000 = 1001 › a2a11. This step decodes the first two symbols and leaves a “remainder” of 1 for the next step. 2. A total of three bits were decoded in the previous step, so the current step XORs the next three bits of C and B. This results in 101 . 010 = 111. Prepending the remainder from the previous step yields 1111, which is decoded to a3 and a remainder of 1. 3. A total of three bits were decoded in the previous step, so the current step XORs the next three bits of C and B, which results in 011 . 111 = 100. Prepending the remainder from the previous step yields 1100 which is decoded to a5 with no remainder. 4. Four bits were decoded in the previous step, so the current step XORs the next four bits of C and B, which results in 1110 . 0011 = 1101, which is decoded to a4. 5. Four bits were decoded in the previous step, so the current step XORs the next four bits of C and B, which results in 1011.1011 = 0000. This indicates the successful end of the decoding procedure. If any data becomes corrupted, the last step will produce something other than L zeros, indicating unsuccessful decoding. Decoding in the reverse direction is identical, except that C is fed to the decoder from end to start (from right to left) to produce substrings of B, which are then decoded and also reversed (to become codewords of B) and sent to the XOR. The first step XORs the reverse of the last four bits of C with the the reverse of the last four bits of B (the four zeros) to produce 1101, which is decoded as a4, reversed, and sent to the XOR. The only overhead is the extra L bits, but L is normally a small number. Figure 4.16 shows the encoder and decoder of this method. forward bitstream B B C B’ B’ offset L bits data symbols codeword VL encoder codeword length reverser + decodable forward bitstream B B C B’ B’ offset L bits data symbols codeword VL encoder codeword length reverser + decodable Encoder Decoder Figure 4.16: XOR-Based Encoding and Decoding. 202 4. Robust VL Codes 4.6 Symmetric Codes A symmetric code is one where every codeword is symmetric. Such a codeword looks the same when read in either direction, which is why symmetric codes are a special case of reversible codes. We can expect a symmetric code to feature higher average length compared with other codes, because the requirement of symmetry restricts the number of available bit patterns of any given length. The material presented here describes one way of selecting a set of symmetric codewords. It is based on [Tsai and Wu 01b], which is an extension of [Takishima et al. 95]. The method starts from a variable-length prefix code, a set of prefix codewords of various lengths, and replaces the codewords with symmetric bit patterns that have the same or similar lengths and also satisfy the prefix property. Figure 4.17 is a good starting point. It shows a complete binary tree with four levels. The symmetric bit patterns at each level are underlined (as an aside, it can be proved by induction that there are 2(i+1)/2 such patterns on level i) and it is clear that, even though the number of symmetric patterns is limited, every path from the root to a leaf passes through several such patterns. It is also clear from the figure that any bit pattern is the prefix of all the patterns below it on the same path. Thus, for example, selecting the pattern 00 implies that we cannot later select any of the symmetric patterns 000, 0000, or any other symmetric pattern below them on the same path. Selecting the 3-bit symmetric pattern 010, on the other hand, does not restrict the choice of 4-bit symmetric patterns. (It will restrict the choice of longer patterns, such as 01010 or 010010, but we are more interested in short patterns.) Thus, a clever algorithm is needed, to select those symmetric patterns at any level that will maximize the number of available symmetric patterns on the levels immediately below them. 0000 000 00 0 1 2 3 1 01 10 11 001 010 011 100 101 110 111 0001 0010 0011 0100 0101 0110 0111 1000 1001 1010 1011 1100 1101 1110 1111 Level Figure 4.17: A 4-Level Full Binary Tree. The analysis presented in [Tsai andWu 01b] proposes the following procedure. Scan the complete binary tree level by level, locating the symmetric bit patterns on each level and ordering them in a special way as shown below. The ordered patterns are then used one by one to replace the original prefix codewords. However, if a symmetric pattern P violates the prefix property (i.e., if a previously-assigned pattern is a prefix of P), then P should be dropped. To order the symmetric patterns of a level, go through the following steps: 1. Ignore the leftmost bit of the pattern. 4.6 Symmetric Codes 203 2. Examine the remaining bits and determine the maximum number M of leastsignificant bits that are still symmetric (the maximum number of symmetric bit suffixes). Table 4.18 lists the orders of the symmetric patterns on levels 3 through 6 of the complete binary tree. In this table, M stands for the maximum number of symmetric bit suffixes and CW is a symmetric pattern. For example, the 5-bit pattern 01110 hasM = 1 because when we ignore its leftmost bit, the remaining four bits 1110 are asymmetric (only the rightmost bit is symmetric). On the other hand, pattern 10101 has M = 3 because the three rightmost bits of 0101 are symmetric. The symmetric bit patterns on each level should be selected and assigned in the order shown in the table (order of increasing M). Thus, if we need 3-bit symmetric patterns, we should first select 010 and 101, and only then 000 and 111. Selecting them in this order maximizes the number of available symmetric patterns on levels 4 and 5. Level 3 Level 4 Level 5 Level 6 M CW M CW M CW M CW 1 010 1 0110 1 01110 1 011110 1 101 1 1001 1 10001 1 100001 2 000 3 0000 2 00100 2 001100 2 111 3 1111 2 11011 2 110011 3 01010 3 010010 3 10101 3 101101 4 00000 5 000000 4 11111 5 111111 Table 4.18: Symmetric Codewords on Levels 3 Through 6, Ordered by Symmetric Bit Suffixes. The following, incomplete, example illustrates the operation of this interesting algorithm. Figure 4.19 shows a Huffman code for the 26 letters of the English alphabet (the individual probabilities of the letters appear in Figure 5.5). The first Huffman codeword is 000. It happens to be symmetric, but Table 4.18 tells us to replace it with 010. It is immediately clear that this is a good choice (as would have been the choice of 101), because Figure 4.17 shows that the path that leads down from 010 to level 4 passes through patterns 0100 and 0101 which are not symmetric. Thus, the choice of 010 does not restrict the future choice of 4-bit symmetric codewords. Figure 4.19 requires six 4-bit symmetric codewords, but there are only four such patterns. They are 0110, 1001, 0000, and 1111, selected in increasing number of symmetric bit suffixes. Pattern 010 is not a prefix of any of these patterns, so they are assigned as the new, symmetric codewords of T, A, O, and N. The remaining two symmetric codewords (of R and I) on this level will have to be longer (five bits each). Figure 4.19 requires 14 5-bit symmetric codewords (and two more, for R and I, are still needed), but there are only eight symmetric 5-bit patterns. Two of them cannot be used because of prefix violation. Pattern 010 is a prefix of 01010 and pattern 0000 is a prefix of 00000. The remaining six 5-bit codewords of Table 4.18 are selected and assigned to the letters R through L, while the letters C through V will obviously have to be assigned longer symmetric codewords. 204 4. Robust VL Codes 000 0010 0011 0100 0101 0110 0111 10000 10001 10010 10011 10100 10101 10110 10111 11000 11001 11010 11011 11100 11101 111100 111101 111110 1111110 1111111 E T A O N R I H S D L C U M F P Y BWG V J K X QZ Figure 4.19: A Huffman Code for the 26-Letter Alphabet. It seems that the symmetric codewords are longer on average than the original Huffman codewords, and this is true in general, because the number of available symmetric codewords of any given length is limited. However, as we go down the tree to longer and longer codewords, the situation improves, because every other level in the complete binary tree doubles the number of available symmetric codewords. Reference [Tsai and Wu 01b] lists Huffman codes and symmetric codes for the 26 letters with average lengths of 4.156 and 4.61 bits per letter, respectively. The special ordering rules above seem arbitrary, but are not difficult to understand intuitively. As an example, we examine the eight symmetric 6-bit patterns. The two patterns of M = 5 become prefixes of patterns at level 7 simply by appending a 0 or a 1. These patterns should therefore be the last ones to be used because they restrict the number of available 7-bit symmetric patterns. The two patterns for M = 3 are better, because they are not the prefixes of any 7-bit or 8-bit symmetric patterns. It is obvious that appending one or two bits to 010010 or 101101 cannot result in a symmetric pattern. Thus, these two patterns are prefixes of 9-bit (or longer) symmetric patterns. When we examine the two patterns for M = 2, the situation is even better. Appending one, two, or even three bits to 100001 or 001100 cannot result in a symmetric pattern. These patterns are the prefixes of 10-bit (or longer) patterns. Similarly, the two 6-bit patterns for M = 1 are the best, because their large asymmetry implies that they can only be prefixes of 11-bit (or longer) patterns. Selecting these two patterns does not restrict the number of symmetric 7-, 8-, 9-, or 10-bit patterns. 4.7 VLEC Codes The simplest approach to robust variable-length codes is to add a parity bit to each codeword. This approach combines compression (source coding) and reliability (channel coding), but keeps the two separate. The two aims are contradictory because the former removes redundancy while the latter adds redundancy. However, any approach to robust compressed data must necessarily reduce compression efficiency. This simple approach has an obvious advantage; it makes it easy to respond to statistical changes in both 4.7 VLEC Codes 205 the source and the channel. If it turns out that the channel is noisy, several parity bits may be appended to each codeword, but the codewords themselves do not have to be modified. If it turns out that certain data symbols occur with higher probability than originally thought, their codewords can be replaced with shorter ones without having to modify the error-control scheme. A different approach to the same problem is to construct a set of variable-length codewords that are sufficiently distant from one another (distant in the sense of Hamming distance). A set of such codewords can be called a variable-length error-correcting (VLEC) code. The downside of this approach is that any changes in source or channel statistics requires a new set of codewords, but the hope is that this approach may result in better overall compression. This approach may be justified if it meets the following two goals: 1. The average length of the codewords turns out to be shorter than the average length of the codewords with parity bits of the previous approach. 2. A decoding algorithm is found that can exploit the large distance between variable-length codewords to detect and even correct errors. Even better, such an algorithm should be able to recover synchronization in cases where bits get corrupted in the communications channel. Several attempts to develop such codes are described here. Alpha-Prompt Codes. The main idea of this technique is to associate each codeword c with a set .(c) of bit patterns of the same length that are at certain Hamming distances from c. If codeword c is transmitted and gets corrupted to c, the decoder checks the entire set .(c) and selects the pattern nearest c. The problem is that the decoder doesn’t know the length of the next codeword, which is why a practical method based on this technique should construct the sets .(c) in a special way. We assume that there are s1 codewords of length L1, s2 codewords of length L2, and so on, up to length m. The set of all the codewords c and all the bit patterns in the individual sets .(c) is constructed as a prefix code (i.e., it satisfies the prefix property and is therefore instantaneous). The decoder starts by reading the next L1 bits from the input. Let’s denote this value by t. If t is a valid codeword c in s1, then t is decoded to c. Otherwise, the decoder checks all the leftmost L1 bits of all the patterns that are L1 bits or longer and compares t to each of them. If t is identical to a pattern in set .(ci), then t is decoded to ci. Otherwise, the decoder again goes over the leftmost L1 bits of all the patterns that are L1 bits or longer and selects the one whose distance from t is minimal. If that pattern is a valid codeword, then t is decoded to it. Otherwise, the decoder reads a few more bits, for a total of L2 bits, and repeats this process for the first L2 bits of all the patterns that are L2 bits long or longer. This complex algorithm (proposed by [Buttigieg 95]) always decodes the input to some codeword if the set of all the codewords c and all the bit patterns in the individual sets .(c) is a prefix code. VLEC Tree. We assume that our code has s1 codewords of length L1, s2 codewords of length L2, and so on, up to length m. The total number of codewords ci is s. As a simple example, consider the code a = 000, b = 0110, and c = 1011 (one 3-bit and two 4-bit codewords). 206 4. Robust VL Codes The VLEC tree method attempts to correct errors by looking at an entire transmission and mapping it to a special VLEC tree. Each node in this tree is connected to s other nodes on lower levels with edges s that are labeled with the codewords ci. Figure 4.20 illustrates an example with the code above. Each path of e edges from the root to level b is therefore associated with a different string of e codewords whose total length is b bits. Thus, node [1] in the figure corresponds to string cba = 1011|0110|000 (11 bits) and node [2] corresponds to the 12-bit string bcc = 0110|1011|1011. There are many possible strings of codewords, but it is also clear that the tree grows exponentially and quickly becomes very wide (the tree in the figure corresponds to strings of up to 12 bits). 12 [1] [2] 11 10 9 8 7 6 4 3 0 1011 0110 000 Figure 4.20: A VLEC Tree. Assuming that the only transmission errors are corrupted bits (i.e., no bits are inserted to or deleted from an encoded string by noise in the communications channel), if an encoded string of b bits is sent, the decoder will receive b bits, some possibly bad. Any b-bit string corresponds to a node on level b, so error-correcting is reduced to the problem of selecting one of the paths that end on that level. The obvious solution is to measure the Hamming distances between the received b-bit string and all the paths that end on level b and select the path with the minimum distance. This simple solution, however, is impractical for the following reasons: Data symbols (and therefore their codewords) have different probabilities, which is why not all b-bit paths are equally likely. If symbol a is very probable, then paths with many occurrences of a are more likely and should be given more weight. Assume that symbol a in our example has probability 0.5, while symbols b and c occur with probability 0.25 each. A probable path such as aabaca should be assigned the weight 4 Χ 0.5 + 2 Χ 0.25 = 2.5, while a less probable path, such as bbccca should be assigned the smaller weight 5 Χ 0.25 + 0.5 = 1.75. When the decoder receives a b-bit string S, it should (1) compute the Hamming distance between S and all the b-bit paths P, (2) divide each distance by the weight of the path, and (3) select the path with the smallest weighted distance. Thus, if two paths with weights 2.5 and 1.75 are at a Hamming distance of 2 from S, the decoder computes 2/2.5 = 0.8 and 2/1.75 = 1.14 and selects the former path. A VLEC tree for even a modest-size alphabet grows too wide very quickly and becomes unmanageable for even short messages. Thus, an algorithm is needed for the 4.7 VLEC Codes 207 decoder to decode an encoded message in short segments. In cases where many errors are rare, a possible approach may be to read the encoded string and encode it normally until the decoder loses synchronization (say, at point A). The decoder then skips bits until it regains synchronization (at point B), and then employs the VLEC tree to the input from A to B. VLEC Trellis. The second point above, of the tree getting too big very quickly, can be overcome by replacing the tree with a trellis structure, as shown in Figure 4.21. S0 S4 S6 S8 S9 S12 S16 S3 S7 S11 S15 S19 S10 S14 S18 S13 S17 a b b b b b c c c c c c c c c c c c c c c c c c c c c c c b b b b b b b b a a Figure 4.21: A VLEC Trellis. The trellis is constructed as follows: 1. Start with state S0. 2. From each state, construct m edges to other states, where each edge corresponds to one of the m codeword lengths Li. Thus, state Si will be connected to states Si+L1 , Si+L2, up to Si+Lm. If any of these states does not exist, it is created when needed. A trellis is a structure, usually made from interwoven pieces of wood, bamboo or metal that supports many types of climbing plant such as sweet peas, grapevines and ivy. —wikipedia.com Notice that the states are arranged in Figure 4.21 in stages, where each stage is a vertical set of states. A comparison of Figures 4.20 and 4.21 shows their similarities, but also shows the important difference between them. Beyond a certain point (in our example, beyond the third stage), the trellis stages stop growing in size regardless of the length of the trellis. It is this difference that makes it possible to develop a reliable decoding algorithm (a modified Viterbi algorithm, described on page 69 of [Buttigieg 95], but not discussed here). An important feature of this algorithm is that it employs a metric (a weight assigned to each trellis edge), and it is possible to define this metric in a way that takes into account the different probabilities of the trellis edges. 208 4. Robust VL Codes 4.7.1 Constructing VLEC Codes p Huffman Takishima Tsai Lakovi΄c E 0.14878 001 001 000 000 T 0.09351 110 110 111 111 A 0.08833 0000 0000 0101 0101 O 0.07245 0100 0100 1010 1010 R 0.06872 0110 1000 0010 0110 N 0.06498 1000 1010 1101 1001 H 0.05831 1010 0101 0100 0011 I 0.05644 1110 11100 1011 1100 S 0.05537 0101 01100 0110 00100 D 0.04376 00010 00010 11001 11011 L 0.04124 10110 10010 10011 01110 U 0.02762 10010 01111 01110 10001 P 0.02575 11110 10111 10001 010010 F 0.02455 01111 11111 001100 101101 M 0.02361 10111 111101 011110 100001 C 0.02081 11111 101101 100001 011110 W 0.01868 000111 000111 1001001 001011 G 0.01521 011100 011101 0011100 110100 Y 0.01521 100110 100111 1100011 0100010 B 0.01267 011101 1001101 0111110 1011101 V 0.01160 100111 01110011 1000001 0100010 K 0.00867 0001100 00011011 00111100 1101011 X 0.00146 00011011 000110011 11000011 10111101 J 0.00080 000110101 0001101011 100101001 010000010 Q 0.00080 0001101001 00011010011 0011101001 0100000010 Z 0.00053 0001101000 000110100011 1001011100 1011111101 Avg. length 4.15572 4.36068 4.30678 4.34534 Table 4.22: Huffman and Three RVLC Codes for the English Alphabet. How can we construct an RVLC with a free distance greater than 1? One such algorithm, proposed by [Lakovi΄c and Villasenor 02] is an extension of an older algorithm due to [Tsai and Wu 01a], which itself is based on the work of [Takishima et al. 95]. This algorithm starts from a set of Huffman codes. It then goes over these codewords level by level, from short to long codewords and replaces some of them with patterns taken from a complete binary tree, similar to what is described in Section 4.6, making sure that the prefix property is not violated while also ascertaining that the free distance of all the codewords that have so far been examined does not drop below the target free distance. If a certain pattern from the binary tree results in a free distance that is too small, that pattern is skipped. Table 4.22 (after [Lakovi΄c and Villasenor 02], see also the very similar Table 4.13) lists a set of Huffman codes for the 26 letters of the English alphabet together with three RVLC codes for the same symbols. The last of these codes has a free distance of 2. These codes were constructed by algorithms proposed by [Takishima et al. 95], [Tsai and Wu 01a], and [Lakovi΄c and Villasenor 02]. 4.7 VLEC Codes 209 Summary. Errors are a fact of life, which is why any practical method for coding, storing, or transmitting data should consider the use of robust codes. The various variable-length codes described in this chapter are robust, which makes them ideal choices for applications where both data compression and data reliability are needed. Good programmers naturally write neat code when left to their own devices. But they also have an array of battle tactics to help write robust code on the front line. —Pete Goodliffe, Code Craft: The Practice of Writing Excellent Code 5 Statistical Methods Statistical data compression methods employ variable-length codes, with the shorter codes assigned to symbols or groups of symbols that appear more often in the data (have a higher probability of occurrence). Designers and implementors of variablelength codes have to deal with the two problems of (1) assigning codes that can be decoded unambiguously and (2) assigning codes with the minimum average size. The first problem is discussed in detail in Chapters 2 through 4, while the second problem is solved in different ways by the methods described here. This chapter is devoted to statistical compression algorithms, such as Shannon- Fano, Huffman, arithmetic coding, and PPM. It is recommended, however, that the reader start with the short presentation of information theory in the Appendix. This presentation covers the principles and important terms used by information theory, especially the terms redundancy and entropy. An understanding of these terms leads to a deeper understanding of statistical data compression, and also makes it possible to calculate how redundancy is reduced, or even eliminated, by the various methods. 5.1 Shannon-Fano Coding Shannon-Fano coding, named after Claude Shannon and Robert Fano, was the first algorithm to construct a set of the best variable-length codes. We start with a set of n symbols with known probabilities (or frequencies) of occurrence. The symbols are first arranged in descending order of their probabilities. The set of symbols is then divided into two subsets that have the same (or almost the same) probabilities. All symbols in one subset get assigned codes that start with a 0, while the codes of the symbols in the other subset start with a 1. Each subset is then recursively divided into two subsubsets of roughly equal probabilities, and the second bit of all the codes is determined in a similar way. When a subset contains just two symbols, DOI 10.1007/978-1-84882-903-9_5, © Springer-Verlag London Limited 2010 D. Salomon, G. Motta, Handbook of Data Compression, 5th ed. 212 5. Statistical Methods their codes are distinguished by adding one more bit to each. The process continues until no more subsets remain. Table 5.1 illustrates the Shannon-Fano algorithm for a seven-symbol alphabet. Notice that the symbols themselves are not shown, only their probabilities. Robert M. Fano was Ford Professor of Engineering, in the Department of Electrical Engineering and Computer Science at the Massachusetts Institute of Technology until his retirement. In 1963 he organized MIT’s Project MAC (now the Computer Science and Artificial Intelligence Laboratory) and was its Director until September 1968. He also served as Associate Head of the Department of Electrical Engineering and Computer Science from 1971 to 1974. Professor Fano chaired the Centennial Study Committee of the Department of Electrical Engineering and Computer Science whose report, “Lifelong Cooperative Education,” was published in October, 1982. Professor Fano was born in Torino, Italy, and did most of his undergraduate work at the School of Engineering of Torino before coming to the United States in 1939. He received the Bachelor of Science degree in 1941 and the Doctor of Science degree in 1947, both in Electrical Engineering from MIT. He has been a member of the MIT staff since 1941 and a member of its faculty since 1947. During World War II, Professor Fano was on the staff of the MIT Radiation Laboratory, working on microwave components and filters. He was also group leader of the Radar Techniques Group of Lincoln Laboratory from 1950 to 1953. He has worked and published at various times in the fields of network theory, microwaves, electromagnetism, information theory, computers and engineering education. He is author of the book entitled Transmission of Information, and co-author of Electromagnetic Fields, Energy and Forces and Electromagnetic Energy Transmission and Radiation. He is also co-author of Volume 9 of the Radiation Laboratory Series. The first step splits the set of seven symbols into two subsets, one with two symbols and a total probability of 0.45 and the other with the remaining five symbols and a total probability of 0.55. The two symbols in the first subset are assigned codes that start with 1, so their final codes are 11 and 10. The second subset is divided, in the second step, into two symbols (with total probability 0.3 and codes that start with 01) and three symbols (with total probability 0.25 and codes that start with 00). Step three divides the last three symbols into 1 (with probability 0.1 and code 001) and 2 (with total probability 0.15 and codes that start with 000). The average size of this code is 0.25Χ2+0.20Χ2+0.15Χ3+0.15Χ3+0.10Χ3+ 0.10 Χ 4 + 0.05 Χ 4 = 2.7 bits/symbol. This is a good result because the entropy (the smallest number of bits needed, on average, to represent each symbol) is - 0.25 log2 0.25 + 0.20 log2 0.20 + 0.15 log2 0.15 + 0.15 log2 0.15 + 0.10 log2 0.10 + 0.10 log2 0.10 + 0.05 log2 0.05 . 2.67. 5.1 Shannon-Fano Coding 213 Prob. Steps Final 1. 0.25 1 1 :11 2. 0.20 1 0 :10 3. 0.15 0 1 1 :011 4. 0.15 0 1 0 :010 5. 0.10 0 0 1 :001 6. 0.10 0 0 0 1 :0001 7. 0.05 0 0 0 0 :0000 Table 5.1: Shannon-Fano Example.  Exercise 5.1: Repeat the calculation above but place the first split between the third and fourth symbols. Calculate the average size of the code and show that it is greater than 2.67 bits/symbol. The code in the table in the answer to Exercise 5.1 has longer average size because the splits, in this case, were not as good as those of Table 5.1. This suggests that the Shannon-Fano method produces better code when the splits are better, i.e., when the two subsets in every split have very close total probabilities. Carrying this argument to its limit suggests that perfect splits yield the best code. Table 5.2 illustrates such a case. The two subsets in every split have identical total probabilities, yielding a code with the minimum average size (zero redundancy). Its average size is 0.25 Χ 2 + 0.25 Χ 2 + 0.125 Χ 3 + 0.125 Χ 3 + 0.125 Χ 3 + 0.125 Χ 3 = 2.5 bits/symbols, which is identical to its entropy. This means that it is the theoretical minimum average size. Prob. Steps Final 1. 0.25 1 1 :11 2. 0.25 1 0 :10 3. 0.125 0 1 1 :011 4. 0.125 0 1 0 :010 5. 0.125 0 0 1 :001 6. 0.125 0 0 0 :000 Table 5.2: Shannon-Fano Balanced Example. The conclusion is that this method produces the best results when the symbols have probabilities of occurrence that are (negative) powers of 2.  Exercise 5.2: Compute the entropy of the codes of Table 5.2. The Shannon-Fano method is easy to implement but the code it produces is generally not as good as that produced by the Huffman method (Section 5.2). 214 5. Statistical Methods 5.2 Huffman Coding David Huffman (1925–1999) Being originally from Ohio, it is no wonder that Huffman went to Ohio State University for his BS (in electrical engineering). What is unusual was his age (18) when he earned it in 1944. After serving in the United States Navy, he went back to Ohio State for an MS degree (1949) and then to MIT, for a PhD (1953, electrical engineering). That same year, Huffman joined the faculty at MIT. In 1967, he made his only career move when he went to the University of California, Santa Cruz as the founding faculty member of the Computer Science Department. During his long tenure at UCSC, Huffman played a major role in the development of the department (he served as chair from 1970 to 1973) and he is known for his motto “my products are my students.” Even after his retirement, in 1994, he remained active in the department, teaching information theory and signal analysis courses. Huffman made significant contributions in several areas, mostly information theory and coding, signal designs for radar and communications, and design procedures for asynchronous logical circuits. Of special interest is the well-known Huffman algorithm for constructing a set of optimal prefix codes for data with known frequencies of occurrence. At a certain point he became interested in the mathematical properties of “zero curvature” surfaces, and developed this interest into techniques for folding paper into unusual sculptured shapes (the so-called computational origami). Huffman coding is a popular method for data compression. It serves as the basis for several popular programs run on various platforms. Some programs use just the Huffman method, while others use it as one step in a multistep compression process. The Huffman method [Huffman 52] is somewhat similar to the Shannon-Fano method. It generally produces better codes, and like the Shannon-Fano method, it produces the best code when the probabilities of the symbols are negative powers of 2. The main difference between the two methods is that Shannon-Fano constructs its codes top to bottom (from the leftmost to the rightmost bits), while Huffman constructs a code tree from the bottom up (builds the codes from right to left). Since its development, in 1952, by D. Huffman, this method has been the subject of intensive research into data compression. Since its development in 1952 by D. Huffman, this method has been the subject of intensive research in data compression. The long discussion in [Gilbert and Moore 59] proves that the Huffman code is a minimum-length code in the sense that no other encoding has a shorter average length. An algebraic approach to constructing the Huffman code is introduced in [Karp 61]. In [Gallager 74], Robert Gallager shows that the redundancy of Huffman coding is at most p1 + 0.086 where p1 is the probability of the most-common symbol in the alphabet. The redundancy is the difference between the average Huffman codeword length and the entropy. Given a large alphabet, such as the 5.2 Huffman Coding 215 set of letters, digits and punctuation marks used by a natural language, the largest symbol probability is typically around 15–20%, bringing the value of the quantity p1+0.086 to around 0.1. This means that Huffman codes are at most 0.1 bit longer (per symbol) than an ideal entropy encoder, such as arithmetic coding. The Huffman algorithm starts by building a list of all the alphabet symbols in descending order of their probabilities. It then constructs a tree, with a symbol at every leaf, from the bottom up. This is done in steps, where at each step the two symbols with smallest probabilities are selected, added to the top of the partial tree, deleted from the list, and replaced with an auxiliary symbol representing the two original symbols. When the list is reduced to just one auxiliary symbol (representing the entire alphabet), the tree is complete. The tree is then traversed to determine the codes of the symbols. This process is best illustrated by an example. Given five symbols with probabilities as shown in Figure 5.3a, they are paired in the following order: 1. a4 is combined with a5 and both are replaced by the combined symbol a45, whose probability is 0.2. 2. There are now four symbols left, a1, with probability 0.4, and a2, a3, and a45, with probabilities 0.2 each. We arbitrarily select a3 and a45, combine them, and replace them with the auxiliary symbol a345, whose probability is 0.4. 3. Three symbols are now left, a1, a2, and a345, with probabilities 0.4, 0.2, and 0.4, respectively. We arbitrarily select a2 and a345, combine them, and replace them with the auxiliary symbol a2345, whose probability is 0.6. 4. Finally, we combine the two remaining symbols, a1 and a2345, and replace them with a12345 with probability 1. The tree is now complete. It is shown in Figure 5.3a “lying on its side” with its root on the right and its five leaves on the left. To assign the codes, we arbitrarily assign a bit of 1 to the top edge, and a bit of 0 to the bottom edge, of every pair of edges. This results in the codes 0, 10, 111, 1101, and 1100. The assignments of bits to the edges is arbitrary. The average size of this code is 0.4 Χ 1 + 0.2 Χ 2 + 0.2 Χ 3 + 0.1 Χ 4 + 0.1 Χ4 = 2.2 bits/symbol, but even more importantly, the Huffman code is not unique. Some of the steps above were chosen arbitrarily, since there were more than two symbols with smallest probabilities. Figure 5.3b shows how the same five symbols can be combined differently to obtain a different Huffman code (11, 01, 00, 101, and 100). The average size of this code is 0.4 Χ 2 + 0.2 Χ 2 + 0.2 Χ 2 + 0.1 Χ 3 + 0.1 Χ3 = 2.2 bits/symbol, the same as the previous code.  Exercise 5.3: Given the eight symbols A, B, C, D, E, F, G, and H with probabilities 1/30, 1/30, 1/30, 2/30, 3/30, 5/30, 5/30, and 12/30, draw three different Huffman trees with heights 5 and 6 for these symbols and calculate the average code size for each tree.  Exercise 5.4: Figure Ans.5d shows another Huffman tree, with height 4, for the eight symbols introduced in Exercise 5.3. Explain why this tree is wrong. It turns out that the arbitrary decisions made in constructing the Huffman tree affect the individual codes but not the average size of the code. Still, we have to answer the obvious question, which of the different Huffman codes for a given set of symbols is best? The answer, while not obvious, is simple: The best code is the one with the 216 5. Statistical Methods 0.4 0.1 0.2 0.2 0.1 0.4 0.1 0.2 0.2 0.1 (a) (b) a3 a345 a4 a45 a5 a2 a2345 a12345 a1 a3 a4 a5 a2 a23 a45 a1 a145 0.2 0.4 0 0 0 0 0 0 0 0 1 1 1 1 0.2 0.4 0.6 1 1 1 1 1.0 0.6 1.0 Figure 5.3: Huffman Codes. smallest variance. The variance of a code measures how much the sizes of the individual codes deviate from the average size (see page 624 for the definition of variance). The variance of code 5.3a is 0.4(1 - 2.2)2 + 0.2(2 - 2.2)2 + 0.2(3 - 2.2)2 + 0.1(4 - 2.2)2 + 0.1(4 - 2.2)2 = 1.36, while the variance of code 5.3b is 0.4(2 - 2.2)2 + 0.2(2 - 2.2)2 + 0.2(2 - 2.2)2 + 0.1(3 - 2.2)2 + 0.1(3 - 2.2)2 = 0.16. Code 5.3b is therefore preferable (see below). A careful look at the two trees shows how to select the one we want. In the tree of Figure 5.3a, symbol a45 is combined with a3, whereas in the tree of 5.3b it is combined with a1. The rule is: When there are more than two smallest-probability nodes, select the ones that are lowest and highest in the tree and combine them. This will combine symbols of low probability with ones of high probability, thereby reducing the total variance of the code. If the encoder simply writes the compressed stream on a file, the variance of the code makes no difference. A small-variance Huffman code is preferable only in cases where the encoder transmits the compressed stream, as it is being generated, over a communications line. In such a case, a code with large variance causes the encoder to generate bits at a rate that varies all the time. Since the bits have to be transmitted at a constant rate, the encoder has to use a buffer. Bits of the compressed stream are entered into the buffer as they are being generated and are moved out of it at a constant rate, to be transmitted. It is easy to see intuitively that a Huffman code with zero variance will enter bits into the buffer at a constant rate, so only a short buffer will be needed. The larger the code variance, the more variable is the rate at which bits enter the buffer, requiring the encoder to use a larger buffer. The following claim is sometimes found in the literature: It can be shown that the size of the Huffman code of a symbol ai with probability Pi is always less than or equal to - log2 Pi. 5.2 Huffman Coding 217 Even though it is correct in many cases, this claim is not true in general. It seems to be a wrong corollary drawn by some authors from the Kraft-MacMillan inequality, Equation (2.3). The authors are indebted to Guy Blelloch for pointing this out and also for the example of Table 5.4.  Exercise 5.5: Find an example where the size of the Huffman code of a symbol ai is greater than - log2 Pi. Pi Code -log2 Pi - log2 Pi .01 000 6.644 7 *.30 001 1.737 2 .34 01 1.556 2 .35 1 1.515 2 Table 5.4: A Huffman Code Example.  Exercise 5.6: It seems that the size of a code must also depend on the number n of symbols (the size of the alphabet). A small alphabet requires just a few codes, so they can all be short; a large alphabet requires many codes, so some must be long. This being so, how can we say that the size of the code of symbol ai depends just on its probability Pi? Figure 5.5 shows a Huffman code for the 26 letters. As a self-exercise, the reader may calculate the average size, entropy, and variance of this code.  Exercise 5.7: Discuss the Huffman codes for equal probabilities. Exercise 5.7 shows that symbols with equal probabilities don’t compress under the Huffman method. This is understandable, since strings of such symbols normally make random text, and random text does not compress. There may be special cases where strings of symbols with equal probabilities are not random and can be compressed. A good example is the string a1a1 . . . a1a2a2 . . . a2a3a3 . . . in which each symbol appears in a long run. This string can be compressed with RLE but not with Huffman codes. Notice that the Huffman method cannot be applied to a two-symbol alphabet. In such an alphabet, one symbol can be assigned the code 0 and the other code 1. The Huffman method cannot assign to any symbol a code shorter than one bit, so it cannot improve on this simple code. If the original data (the source) consists of individual bits, such as in the case of a bi-level (monochromatic) image, it is possible to combine several bits (perhaps four or eight) into a new symbol and pretend that the alphabet consists of these (16 or 256) symbols. The problem with this approach is that the original binary data may have certain statistical correlations between the bits, and some of these correlations would be lost when the bits are combined into symbols. When a typical bi-level image (a painting or a diagram) is digitized by scan lines, a pixel is more likely to be followed by an identical pixel than by the opposite one. We therefore have a file that can start with either a 0 or a 1 (each has 0.5 probability of being the first bit). A zero is more likely to be followed by another 0 and a 1 by another 1. Figure 5.6 is a finite-state 218 5. Statistical Methods 000 E .1300 0010 T .0900 0011 A .0800 0100 O .0800 0101 N .0700 0110 R .0650 0111 I .0650 10000 H .0600 10001 S .0600 10010 D .0400 10011 L .0350 10100 C .0300 10101 U .0300 10110 M .0300 10111 F .0200 11000 P .0200 11001 Y .0200 11010 B .0150 11011 W .0150 11100 G .0150 11101 V .0100 111100 J .0050 111101 K .0050 111110 X .0050 1111110 Q .0025 1111111 Z .0025 .005 .11 .010 .010 .020 .025 .045 .070 .115 .305 .420 .580 .30 .28 .195 1.0 1 1 0 0 1 0 1 0 1 0 0 1 Figure 5.5: A Huffman Code for the 26-Letter Alphabet. machine illustrating this situation. If these bits are combined into, say, groups of eight, the bits inside a group will still be correlated, but the groups themselves will not be correlated by the original pixel probabilities. If the input stream contains, e.g., the two adjacent groups 00011100 and 00001110, they will be encoded independently, ignoring the correlation between the last 0 of the first group and the first 0 of the next group. Selecting larger groups improves this situation but increases the number of groups, which implies more storage for the code table and longer time to calculate the table.  Exercise 5.8: How does the number of groups increase when the group size increases from s bits to s + n bits? A more complex approach to image compression by Huffman coding is to create several complete sets of Huffman codes. If the group size is, e.g., eight bits, then several sets of 256 codes are generated. When a symbol S is to be encoded, one of the sets is selected, and S is encoded using its code in that set. The choice of set depends on the 5.2 Huffman Coding 219 0 1 s 0,50% 1,50% 0,40% 1,60% 1,33% 0,67% Start Figure 5.6: A Finite-State Machine. symbol preceding S.  Exercise 5.9: Imagine an image with 8-bit pixels where half the pixels have values 127 and the other half have values 128. Analyze the performance of RLE on the individual bitplanes of such an image, and compare it with what can be achieved with Huffman coding. 5.2.1 Dual Tree Coding Dual tree coding, an idea due to G. H. Freeman ([Freeman 91] and [Freeman 93]), combines Tunstall and Huffman coding in an attempt to improve the latter’s performance for a 2-symbol alphabet. The idea is to use the Tunstall algorithm to extend such an alphabet from 2 symbols to 2k strings of symbols, and select k such that the probabilities of the strings will be close to negative powers of 2. Once this is achieved, the strings are assigned Huffman codes and the input stream is compressed by replacing the strings with these codes. This approach is illustrated here by a simple example. Given a binary source that emits two symbols a and b with probabilities 0.15 and 0.85, respectively, we try to compress it in four different ways as follows: 1. We apply the Huffman algorithm directly to the two symbols. This simply assigns the two 1-bit codes 0 and 1 to a and b, so there is no compression. 2. We combine the two symbols to obtain the four 2-symbol strings aa, ab, ba, and bb, with probabilities 0.0225, 0.1275, 0.1275, and 0.7225, respectively. The four strings are assigned Huffman codes as shown in Figure 5.7a, and it is obvious that the average code length is 0.0225 Χ 3 + 0.1275 Χ 3 + 0.1275 Χ 2 + 0.7225 Χ 1 = 1.4275 bits. On average, each 2-symbol string is compressed to 1.4275 bits, yielding a compression ratio of 1.4275/2 . 0.714. 3. We apply Tunstall’s algorithm to obtain the four strings bbb, bba, ba, and a with probabilities 0.614, 0.1084, 0.1275, and 0.15, respectively. The resulting parse tree is shown in Figure 5.7b. Tunstall’s method compresses these strings by replacing each with a 2-bit code. Given a string of 257 bits with these probabilities, we expect the strings bbb, bba, ba, and a to occur 61, 11, 13, and 15 times, respectively, for a total of 100 strings. Thus, Tunstall’s method compresses the 257 input bits to 2 Χ 100 = 200 bits, for a compression ratio of 200/257 . 0.778. 4. We now change the probabilities of the four strings above to negative powers of 2, because these are the best values for the Huffman method. Strings bbb, bba, 220 5. Statistical Methods (a) (b) (c) 0 0 0 1 1 1 a a b ba ba bb bba bba bbb bbb aa ab ba bb .15 .25 .85 .1275 .7225 .50 .1084 .614 Figure 5.7: Dual Tree Coding. ba, and a are thus assigned the probabilities 0.5, 0.125, 0.125, and 0.25, respectively. The resulting Huffman code tree is shown in Figure 5.7c and it is easy to see that the 61, 11, 13, and 15 occurrences of these strings will be compressed to a total of 61 Χ 1 + 11 Χ 3 + 13 Χ 3 + 15 Χ 2 = 163 bits, resulting in a compression ratio of 163/257 . 0.634, much better. To summarize, applying the Huffman method to a 2-symbol alphabet produces no compression. Combining the individual symbols in strings as in 2 above or applying the Tunstall method as in 3, produce moderate compression. In contrast, combining the strings produced by Tunstall with the codes generated by the Huffman method, results in much better performance. The dual tree method starts by constructing the Tunstall parse tree and then using its leaf nodes to construct a Huffman code tree. The only (still unsolved) problem is determining the best value of k. In our example, we iterated the Tunstall algorithm until we had 22 = 4 strings, but iterating more times may have resulted in strings whose probabilities are closer to negative powers of 2. 5.2.2 Huffman Decoding Before starting the compression of a data stream, the compressor (encoder) has to determine the codes. It does that based on the probabilities (or frequencies of occurrence) of the symbols. The probabilities or frequencies have to be written, as side information, on the compressed stream, so that any Huffman decompressor (decoder) will be able to decompress the stream. This is easy, since the frequencies are integers and the probabilities can be written as scaled integers. It normally adds just a few hundred bytes to the compressed stream. It is also possible to write the variable-length codes themselves on the stream, but this may be awkward, because the codes have different sizes. It is also possible to write the Huffman tree on the stream, but this may require more space than just the frequencies. In any case, the decoder must know what is at the start of the stream, read it, and construct the Huffman tree for the alphabet. Only then can it read and decode the rest of the stream. The algorithm for decoding is simple. Start at the root and read the first bit off the compressed stream. If it is zero, follow the bottom edge of the tree; if it is one, follow the top edge. Read the next bit and move another edge toward the leaves of the tree. When the decoder gets to a leaf, it finds the original, uncompressed code of the symbol (normally its ASCII code), and that code is emitted by the decoder. The 5.2 Huffman Coding 221 process starts again at the root with the next bit. This process is illustrated for the five-symbol alphabet of Figure 5.8. The foursymbol input string a4a2a5a1 is encoded into 1001100111. The decoder starts at the root, reads the first bit 1, and goes up. The second bit 0 sends it down, as does the third bit. This brings the decoder to leaf a4, which it emits. It again returns to the root, reads 110, moves up, up, and down, to reach leaf a2, and so on. 1 2 3 4 5 1 1 0 0 Figure 5.8: Huffman Codes for Equal Probabilities. 5.2.3 Fast Huffman Decoding Decoding a Huffman-compressed file by sliding down the code tree for each symbol is conceptually simple, but slow. The compressed file has to be read bit by bit and the decoder has to advance a node in the code tree for each bit. The method of this section, originally conceived by [Choueka et al. 85] but later reinvented by others, uses preset partial-decoding tables. These tables depend on the particular Huffman code used, but not on the data to be decoded. The compressed file is read in chunks of k bits each (where k is normally 8 or 16 but can have other values) and the current chunk is used as a pointer to a table. The table entry that is selected in this way can decode several symbols and it also points the decoder to the table to be used for the next chunk. As an example, consider the Huffman code of Figure 5.3a, where the five codewords are 0, 10, 111, 1101, and 1100. The string of symbols a1a1a2a4a3a1a5 . . . is compressed by this code to the string 0|0|10|1101|111|0|1100 . . .. We select k = 3 and read this string in 3-bit chunks 001|011|011|110|110|0 . . .. Examining the first chunk, it is easy to see that it should be decoded into a1a1 followed by the single bit 1 which is the prefix of another codeword. The first chunk is 001 = 110, so we set entry 1 of the first table (table 0) to the pair (a1a1, 1). When chunk 001 is used as a pointer to table 0, it points to entry 1, which immediately provides the decoder with the two decoded symbols a1a1 and also directs it to use table 1 for the next chunk. Table 1 is used when a partially-decoded chunk ends with the single-bit prefix 1. The next chunk is 011 = 310, so entry 3 of table 1 corresponds to the encoded bits 1|011. Again, it is easy to see that these should be decoded to a2 and there is the prefix 11 left over. Thus, entry 3 of table 1 should be (a2, 2). It provides the decoder with the single symbol a2 and also directs it to use table 2 next (the table that corresponds to prefix 11). The next chunk is again 011 = 310, so entry 3 of table 2 corresponds to the encoded bits 11|011. It is again obvious that these 222 5. Statistical Methods i‹0; output‹null; repeat j‹input next chunk; (s,i)‹Tablei[j]; append s to output; until end-of-input Figure 5.9: Fast Huffman Decoding. should be decoded to a4 with a prefix of 1 left over. This process continues until the end of the encoded input. Figure 5.9 is the simple decoding algorithm in pseudocode. Table 5.10 lists the four tables required to decode this code. It is easy to see that they correspond to the prefixes . (null), 1, 11, and 110. A quick glance at Figure 5.3a shows that these correspond to the root and the four interior nodes of the Huffman code tree. Thus, each partial-decoding table corresponds to one of the four prefixes of this code. The number m of partial-decoding tables therefore equals the number of interior nodes (plus the root) which is one less than the number N of symbols of the alphabet. T0 = . T1 = 1 T2 = 11 T3 = 110 000 a1a1a1 0 1|000 a2a1a1 0 11|000 a5a1 0 110|000 a5a1a1 0 001 a1a1 1 1|001 a2a1 1 11|001 a5 1 110|001 a5a1 1 010 a1a2 0 1|010 a2a2 0 11|010 a4a1 0 110|010 a5a2 0 011 a1 2 1|011 a2 2 11|011 a4 1 110|011 a5 2 100 a2a1 0 1|100 a5 0 11|100 a3a1a1 0 110|100 a4a1a1 0 101 a2 1 1|101 a4 0 11|101 a3a1 1 110|101 a4a1 1 110 - 3 1|110 a3a1 0 11|110 a3a2 0 110|110 a4a2 0 111 a3 0 1|111 a3 1 11|111 a3 2 110|111 a4 2 Table 5.10: Partial-Decoding Tables for a Huffman Code. Notice that some chunks (such as entry 110 of table 0) simply send the decoder to another table and do not provide any decoded symbols. Also, there is a tradeoff between chunk size (and thus table size) and decoding speed. Large chunks speed up decoding, but require large tables. A large alphabet (such as the 128 ASCII characters or the 256 8-bit bytes) also requires a large set of tables. The problem with large tables is that the decoder has to set up the tables after it has read the Huffman codes from the compressed stream and before decoding can start, and this process may preempt any gains in decoding speed provided by the tables. To set up the first table (table 0, which corresponds to the null prefix .), the decoder generates the 2k bit patterns 0 through 2k - 1 (the first column of Table 5.10) and employs the decoding method of Section 5.2.2 to decode each pattern. This yields the second column of Table 5.10. Any remainders left are prefixes and are converted by the decoder to table numbers. They become the third column of the table. If no remainder is left, the third column is set to 0 (use table 0 for the next chunk). Each of the other partial-decoding tables is set in a similar way. Once the decoder decides that 5.2 Huffman Coding 223 table 1 corresponds to prefix p, it generates the 2k patterns p|00 . . . 0 through p|11 . . . 1 that become the first column of that table. It then decodes that column to generate the remaining two columns. This method was conceived in 1985, when storage costs were considerably higher than today (early 2007). This prompted the developers of the method to find ways to cut down the number of partial-decoding tables, but these techniques are less important today and are not described here. Truth is stranger than fiction, but this is because fiction is obliged to stick to probability; truth is not. —Anonymous 5.2.4 Average Code Size Figure 5.13a shows a set of five symbols with their probabilities and a typical Huffman tree. Symbol A appears 55% of the time and is assigned a 1-bit code, so it contributes 0.55 ·1 bits to the average code size. Symbol E appears only 2% of the time and is assigned a 4-bit Huffman code, so it contributes 0.02·4 = 0.08 bits to the code size. The average code size is therefore calculated to be 0.55 · 1 + 0.25 · 2 + 0.15 · 3 + 0.03 · 4 + 0.02 · 4 = 1.7 bits per symbol. Surprisingly, the same result is obtained by adding the values of the four internal nodes of the Huffman code-tree 0.05 + 0.2 + 0.45 + 1 = 1.7. This provides a way to calculate the average code size of a set of Huffman codes without any multiplications. Simply add the values of all the internal nodes of the tree. Table 5.11 illustrates why this works. .05 = .02+ .03 .20 =.05+.15 =.02+ .03+ .15 .45 =.20+.25 =.02+ .03+ .15+ .25 1 .0 =.45+.55 =.02+ .03+ .15+ .25+ .55 Table 5.11: Composition of Nodes. 0 .05= =0.02 + 0.03 + · · · a1 =0 .05 + . . .= 0.02 + 0.03 + · · · a2 =a1 + . . .= 0.02 + 0.03 + · · · .. . = ad-2 =ad-3 + . . .= 0.02 + 0.03 + · · · 1 .0 =ad-2 + . . .= 0.02 + 0.03 + · · · Table 5.12: Composition of Nodes. (Internal nodes are shown in italics in this table.) The left column consists of the values of all the internal nodes. The right columns show how each internal node is the sum of some of the leaf nodes. Summing the values in the left column yields 1.7, and summing the other columns shows that this 1.7 is the sum of the four values 0.02, the four values 0.03, the three values 0.15, the two values 0.25, and the single value 0.55. This argument can be extended to the general case. It is easy to show that, in a Huffman-like tree (a tree where each node is the sum of its children), the weighted sum 224 5. Statistical Methods A 0.55 B 0.25 C 0.15 D 0.03 E 0.02 0.05 0.2 0.45 1 0.02 0.03 0.05 1 d (b) (a) ad-2 a1 Figure 5.13: Huffman Code-Trees. of the leaves, where the weights are the distances of the leaves from the root, equals the sum of the internal nodes. (This property has been communicated to us by John M. Motil.) Figure 5.13b shows such a tree, where we assume that the two leaves 0.02 and 0.03 have d-bit Huffman codes. Inside the tree, these leaves become the children of internal node 0.05, which, in turn, is connected to the root by means of the d-2 internal nodes a1 through ad-2. Table 5.12 has d rows and shows that the two values 0.02 and 0.03 are included in the various internal nodes exactly d times. Adding the values of all the internal nodes produces a sum that includes the contributions 0.02 · d + 0.03 · d from the two leaves. Since these leaves are arbitrary, it is clear that this sum includes similar contributions from all the other leaves, so this sum is the average code size. Since this sum also equals the sum of the left column, which is the sum of the internal nodes, it is clear that the sum of the internal nodes equals the average code size. Notice that this proof does not assume that the tree is binary. The property illustrated here exists for any tree where a node contains the sum of its children. 5.2 Huffman Coding 225 5.2.5 Number of Codes Since the Huffman code is not unique, the natural question is: How many different codes are there? Figure 5.14a shows a Huffman code-tree for six symbols, from which we can answer this question in two different ways. Answer 1. The tree of 5.14a has five interior nodes, and in general, a Huffman codetree for n symbols has n-1 interior nodes. Each interior node has two edges coming out of it, labeled 0 and 1. Swapping the two labels produces a different Huffman code-tree, so the total number of different Huffman code-trees is 2n-1 (in our example, 25 or 32). The tree of Figure 5.14b, for example, shows the result of swapping the labels of the two edges of the root. Table 5.15a,b lists the codes generated by the two trees. 1 2 3 4 5 6 .11 .12 .13 .14 .24 .26 .11 .12 .13 .14 .24 .26 0 1 0 0 0 0 1 1 1 1 1 2 3 4 5 6 0 1 0 0 0 0 1 1 1 1 (a) (b) 000 100 000 001 101 001 100 000 010 101 001 011 01 11 10 11 01 11 (a) (b) (c) Figure 5.14: Two Huffman Code-Trees. Table 5.15. Answer 2. The six codes of Table 5.15a can be divided into the four classes 00x, 10y, 01, and 11, where x and y are 1-bit each. It is possible to create different Huffman codes by changing the first two bits of each class. Since there are four classes, this is the same as creating all the permutations of four objects, something that can be done in 4! = 24 ways. In each of the 24 permutations it is also possible to change the values of x and y in four different ways (since they are bits) so the total number of different Huffman codes in our six-symbol example is 24 Χ 4 = 96. The two answers are different because they count different things. Answer 1 counts the number of different Huffman code-trees, while answer 2 counts the number of different Huffman codes. It turns out that our example can generate 32 different code-trees but only 94 different codes instead of 96. This shows that there are Huffman codes that cannot be generated by the Huffman method! Table 5.15c shows such an example. A look at the trees of Figure 5.14 should convince the reader that the codes of symbols 5 and 6 must start with different bits, but in the code of Table 5.15c they both start with 1. This code is therefore impossible to generate by any relabeling of the nodes of the trees of Figure 5.14. 5.2.6 Ternary Huffman Codes The Huffman code is not unique. Moreover, it does not have to be binary! The Huffman method can easily be applied to codes based on other number systems. Figure 5.16a 226 5. Statistical Methods shows a Huffman code tree for five symbols with probabilities 0.15, 0.15, 0.2, 0.25, and 0.25. The average code size is 2Χ0.25 + 3Χ0.15 + 3Χ0.15 + 2Χ0.20 + 2Χ0.25 = 2.3 bits/symbol. Figure 5.16b shows a ternary Huffman code tree for the same five symbols. The tree is constructed by selecting, at each step, three symbols with the smallest probabilities and merging them into one parent symbol, with the combined probability. The average code size of this tree is 2Χ0.15 + 2Χ0.15 + 2Χ0.20 + 1Χ0.25 + 1Χ0.25 = 1.5 trits/symbol. Notice that the ternary codes use the digits 0, 1, and 2.  Exercise 5.10: Given seven symbols with probabilities .02, .03, .04, .04, .12, .26, and .49, we construct binary and ternary Huffman code-trees for them and calculate the average code size in each case. 5.2.7 Height Of A Huffman Tree The height of the code-tree generated by the Huffman algorithm may sometimes be important because the height is also the length of the longest code in the tree. The Deflate method (Section 6.25), for example, limits the lengths of certain Huffman codes to just three bits. It is easy to see that the shortest Huffman tree is created when the symbols have equal probabilities. If the symbols are denoted by A, B, C, and so on, then the algorithm combines pairs of symbols, such A and B, C and D, in the lowest level, and the rest of the tree consists of interior nodes as shown in Figure 5.17a. The tree is balanced or close to balanced and its height is log2 n. In the special case where the number of symbols n is a power of 2, the height is exactly log2 n. In order to generate the tallest tree, we need to assign probabilities to the symbols such that each step in the Huffman method will increase the height of the tree by 1. Recall that each step in the Huffman algorithm combines two symbols. Thus, the tallest tree is obtained when the first step combines two of the n symbols and each subsequent step combines the result of its predecessor with one of the remaining symbols (Figure 5.17b). The height of the complete tree is therefore n - 1, and it is referred to as a lopsided or unbalanced tree. It is easy to see what symbol probabilities result in such a tree. Denote the two smallest probabilities by a and b. They are combined in the first step to form a node whose probability is a + b. The second step will combine this node with an original symbol if one of the symbols has probability a + b (or smaller) and all the remaining symbols have greater probabilities. Thus, after the second step, the root of the tree has probability a + b + (a + b) and the third step will combine this root with one of the remaining symbols if its probability is a + b + (a + b) and the probabilities of the remaining n - 4 symbols are greater. It does not take much to realize that the symbols have to have probabilities p1 = a, p2 = b, p3 = a+b = p1+p2, p4 = b+(a+b) = p2+p3, p5 = (a + b) + (a + 2b) = p3 + p4, p6 = (a + 2b) + (2a + 3b) = p4 + p5, and so on (Figure 5.17c). These probabilities form a Fibonacci sequence whose first two elements 5.2 Huffman Coding 227 (a) .15 .15 .20 .15 .15 .20 .50 .25 .25 .25 .25 .30 .45 .55 1.0 1.0 (b) (c) (d) .02 .03 .04 .02 .03 .04 .09 .04 .12 .26 .25 .49 .04 .08 .13 .12 .26 .25 .49 .51 1.0 1.0 .05 Figure 5.16: Binary and Ternary Huffman Code-Trees. are a and b. As an example, we select a = 5 and b = 2 and generate the 5-number Fibonacci sequence 5, 2, 7, 9, and 16. These five numbers add up to 39, so dividing them by 39 produces the five probabilities 5/39, 2/39, 7/39, 9/39, and 15/39. The Huffman tree generated by them has a maximal height (which is 4). In principle, symbols in a set can have any probabilities, but in practice, the probabilities of symbols in an input file are computed by counting the number of occurrences of each symbol. Imagine a text file where only the nine symbols A through I appear. In order for such a file to produce the tallest Huffman tree, where the codes will have lengths from 1 to 8 bits, the frequencies of occurrence of the nine symbols have to form a Fibonacci sequence of probabilities. This happens when the frequencies of the symbols are 1, 1, 2, 3, 5, 8, 13, 21, and 34 (or integer multiples of these). The sum of these 228 5. Statistical Methods 000 001 010 011 100 101 110 111 (a) (b) (c) a+b a+2b 2a+3b 3a+5b 5a+8b a b 0 10 110 1110 11110 11111 Figure 5.17: Shortest and Tallest Huffman Trees. frequencies is 88, so our file has to be at least that long in order for a symbol to have 8-bit Huffman codes. Similarly, if we want to limit the sizes of the Huffman codes of a set of n symbols to 16 bits, we need to count frequencies of at least 4180 symbols. To limit the code sizes to 32 bits, the minimum data size is 9,227,464 symbols. If a set of symbols happens to have the Fibonacci probabilities and therefore results in a maximal-height Huffman tree with codes that are too long, the tree can be reshaped (and the maximum code length shortened) by slightly modifying the symbol probabilities, so they are not much different from the original, but do not form a Fibonacci sequence. 5.2.8 Canonical Huffman Codes The code of Table 5.15c has a simple interpretation. It assigns the first four symbols the 3-bit codes 0, 1, 2, 3, and the last two symbols the 2-bit codes 2 and 3. This is an example of a canonical Huffman code. The word “canonical” means that this particular code has been selected from among the several (or even many) possible Huffman codes because its properties make it easy and fast to use. Table 5.18 shows a slightly bigger example of a canonical Huffman code. Imagine a set of 16 symbols (whose probabilities are irrelevant and are not shown) such that four symbols are assigned 3-bit codes, five symbols are assigned 5-bit codes, and the remaining seven symbols are assigned 6-bit codes. Table 5.18a shows a set of possible Huffman codes, while Table 5.18b shows a set of canonical Huffman codes. It is easy to see that the seven 6-bit canonical codes are simply the 6-bit integers 0 through 6. The five codes are the 5-bit integers 4 through 8, and the four codes are the 3-bit integers 3 through 6. We first show how these codes are generated and then how they are used. The top row (length) of Table 5.19 lists the possible code lengths, from 1 to 6 bits. The second row (numl) lists the number of codes of each length, and the bottom row (first) lists the first code in each group. This is why the three groups of codes start with values 3, 4, and 0. To obtain the top two rows we need to compute the lengths of all the Huffman codes for the given alphabet (see below). The third row is computed by setting first[6]:=0; and iterating for l:=5 downto 1 do first[l]:=(first[l+1]+numl[l+1])/2; This guarantees that all the 3-bit prefixes of codes longer than three bits will be less than first[3] (which is 3), all the 5-bit prefixes of codes longer than five bits will be less than first[5] (which is 4), and so on. 5.2 Huffman Coding 229 1: 000 011 9: 10100 01000 2: 001 100 10: 101010 000000 3: 010 101 11: 101011 000001 4: 011 110 12: 101100 000010 5: 10000 00100 13: 101101 000011 6: 10001 00101 14: 101110 000100 7: 10010 00110 15: 101111 000101 8: 10011 00111 16: 110000 000110 (a) (b) (a) (b) length: 1 2 3 4 5 6 numl: 0 0 4 0 5 7 first: 2 4 3 5 4 0 Table 5.18. Table 5.19. Now for the use of these unusual codes. Canonical Huffman codes are useful in cases where the alphabet is large and where fast decoding is mandatory. Because of the way the codes are constructed, it is easy for the decoder to identify the length of a code by reading and examining input bits one by one. Once the length is known, the symbol can be found in one step. The pseudocode listed here shows the rules for decoding: l:=1; input v; while v F[b], entry a has to be assigned a short code, and this is done by swapping pointers P[a] and P[b] (and also the corresponding Q pointers). 5.4 MNP5 243 0 F P Q C 0 0 0 0 0 0 10 0 0 F P Q C 0 0 0 0 0 0 10 0 0 F P Q C 0 0 0 0 0 0 1 0 0 (a) (b) (c) a b c a a d Figure 5.26: Swapping Pointers in the MNP5 Code Table (Part I). (b) The same process is repeated. The algorithm again starts with pointer P[a], which now points higher, to entry b. Assuming that P[a] contains the index k, the algorithm examines pointer Q[k-1], which points to entry c. Since F[a] > F[c], entry a should be assigned a code shorter than that of c. This again is done by swapping pointers, this time P[a] and P[c]. (c) This process is repeated, and since F[a] is greater than all the frequencies above it, pointers are swapped until P[a] points to the top entry, d. At this point entry a has been assigned the shortest code. (d) We now assume that the next data item has been input, and F[m] incremented to 1. Pointers P[m] and the one above it are swapped as in (a) above. (e) After a few more swaps, P[m] is now pointing to entry n (the one just below a). The next step performs j:=P[m]; b:=Q[j-1], and the algorithm compares F[m] to F[b]. Since F[m] > F[b], pointers are swapped as shown in Figure 5.26f. (f) After the swap, P[m] is pointing to entry a and P[b] is pointing to entry n. (g) After a few more swaps, pointer P[m] points to the second entry p. This is how entry m is assigned the second-shortest code. Pointer P[m] is not swapped with P[a], since they have the same frequencies. (h) We now assume that the third data item has been input and F[x] incremented. Pointers P[x] and P[y] are swapped. (i) After some more swaps, pointer P[x] points to the third table entry z. This is how entry x is assigned the third shortest code. Assuming that F[x] is incremented next, the reader is invited to try to figure out how P[x] is swapped, first with P[m] and then with P[a], so that entry x is assigned the shortest code. 244 5. Statistical Methods 0 0 0 0 0 0 0 1 0 1 0 0 0 0 0 1 01 0 1 0 0 0 0 01 01 0 1 (g) (h) (i) d p m x y x z 0 0 0 0 0 0 0 1 0 1 0 0 0 0 0 0 0 1 0 1 0 0 0 0 0 0 01 0 1 (d) (e) (f) a m m n m b a b Figure 5.26 (Continued). 5.5 MNP7 245 F[i]:=F[i]+1; repeat forever j:=P[i]; if j=1 then exit; j:=Q[j-1]; if F[i]<=F[j] then exit else tmp:=P[i]; P[i]:=P[j]; P[j]:=tmp; tmp:=Q[P[i]]; Q[P[i]]:=Q[P[j]]; Q[P[j]]:=tmp endif; end repeat Figure 5.27: Swapping Pointers in MNP5. The pseudo-code of Figure 5.27 summarizes the pointer swapping process. Are no probabilities to be accepted, merely because they are not certainties? —Jane Austen, Sense and Sensibility 5.5 MNP7 More complex and sophisticated than MNP5, MNP7 combines run-length encoding with a two-dimensional variant of adaptive Huffman coding. Stage 1 identifies runs and emits three copies of the run character, followed by a 4-bit count of the remaining characters in the run. A count of zero implies a run of length 3, and a count of 15 (the largest possible in a 4-bit nibble), a run of length 18. Stage 2 starts by assigning to each character a complete table with many variable-length codes. When a character C is read, one of the codes in its table is selected and output, depending on the character preceding C in the input stream. If this character is, say, P, then the frequency count of the pair (digram) PC is incremented by 1, and table rows may be swapped, using the same algorithm as for MNP5, to move the pair to a position in the table that has a shorter code. MNP7 is therefore based on a first-order Markov model, where each item is processed depending on the item and one of its predecessors. In a k-order Markov model, an item is processed depending on itself and k of its predecessors (not necessarily the k immediate ones). Here are the details. Each of the 256 8-bit bytes gets a table of codes assigned, of size 256 Χ 2, where each row corresponds to one of the 256 possible bytes. Column 1 of the table is initialized to the integers 0 through 255, and column 2 (the frequency counts) is initialized to all zeros. The result is 256 tables, each a double column of 256 rows (Table 5.28a). Variable-length codes are assigned to the rows, such that the first code is 1-bit long, and the others get longer towards the bottom of the table. The codes are stored in an additional code table that never changes. When a character C is input (the current character to be compressed), its value is used as a pointer, to select one of the 256 tables. The first column of the table is 246 5. Statistical Methods Preced. Char. Current character 0 1 2 . . . 254 255 0 0 0 0 0 0 . . . 0 0 0 0 1 0 1 0 1 0 . . . 1 0 1 0 2 0 2 0 2 0 . . . 2 0 2 0 3 0 3 0 3 0 . . . 3 0 3 0 ... ... ... ... ... 254 0 254 0 254 0 . . . 254 0 254 0 255 0 255 0 255 0 . . . 255 0 255 0 (a) . . .a b c d e. . . t l h o d h e o a r c u r e s ... ... ... ... ... (b) Table 5.28: The MNP7 Code Tables. searched, to find the row with the 8-bit value of the preceding character P. Once the row is found, the code from the same row in the code table is emitted and the count in the second column is incremented by 1. Rows in the table may be swapped if the new count of the digram PC is large enough. After enough characters have been input and rows swapped, the tables start reflecting the true digram frequencies of the data. Table 5.28b shows a possible state assuming that the digrams ta, ha, ca, lb, eb, ub, hc, etc., are common. Since the top digram is encoded into 1 bit, MNP7 can be very efficient. If the original data consists of text in a natural language, where certain digrams are very common, MNP7 normally produces a high compression ratio. His only other visitor was a woman: the woman who had attended his reading. At the time she had seemed to him to be the only person present who had paid the slightest attention to his words. With kittenish timidity she approached his table. Richard bade her welcome, and meant it, and went on meaning it as she extracted from her shoulder pouch a copy of a novel written not by Richard Tull but by Fyodor Dostoevsky. The Idiot. Standing beside him, leaning over him, her face awfully warm and near, she began to leaf through its pages, explaining. The book too was stained, not by gouts of blood but by the vying colors of two highlighting pens, one blue, one pink. And not just two pages but the whole six hundred. Every time the letters h and e appeared together, as in the, then, there, as in forehead, Pashlishtchev, sheepskin, they were shaded in blue. Every time the letters s, h, and e appeared together, as in she, sheer, ashen, sheepskin, etc., they were shaded in pink. And since every she contained a he, the predominance was unarguably and unsurprisingly masculine. Which was exactly her point. “You see?” she said with her hot breath, breath redolent of metallic medications, of batteries and printing-plates. “You see?”. . . The organizers knew all about this woman—this unfortunate recurrence, this indefatigable drag—and kept coming over to try and coax her away. —Martin Amis, The Information 5.6 Reliability 247 5.6 Reliability The chief downside of variable-length codes is their vulnerability to errors. The prefix property is used to decode those codes, so an error in a single bit can cause the decompressor to lose synchronization and be unable to decode the rest of the compressed stream. In the worst case, the decompressor may even read, decode, and interpret the rest of the compressed data incorrectly, without realizing that a problem has occurred. Example: Using the code of Figure 5.5 the string CARE is coded into 10100 0011 0110 000 (without the spaces). Assuming the error 10 0 00 0011 0110 000, the decompressor will not notice any problem but will decode the string as HARE.  Exercise 5.17: What will happen in the case 11 1 11 0011 0110 000 . . . (the string WARE . . . with one bad bit)? Users of Huffman codes have noticed long ago that these codes recover quickly from an error. However, if Huffman codes are used to code run lengths, then this property does not help, since all runs would be shifted after an error. A simple way of adding reliability to variable-length codes is to break a long compressed stream, as it is being transmitted, into groups of 7 bits and add a parity bit to each group. This way, the decompressor will at least be able to detect a problem and output an error message or ask for a retransmission. It is, of course, possible to add more than one parity bit to a group of data bits, thereby making it more reliable. However, reliability is, in a sense, the opposite of compression. Compression is done by decreasing redundancy, while reliability is achieved by increasing it. The more reliable a piece of data is, the less compressed it is, so care should be taken when the two operations are used together. Some Important Standards Groups and Organizations ANSI (American National Standards Institute) is the private sector voluntary standardization system for the United States. Its members are professional societies, consumer groups, trade associations, and some government regulatory agencies (it is a federation). It collects and distributes standards developed by its members. Its mission is the enhancement of global competitiveness of U.S. business and the American quality of life by promoting and facilitating voluntary consensus standards and conformity assessment systems and promoting their integrity. ANSI was founded in 1918 by five engineering societies and three government agencies, and it remains a private, nonprofit membership organization whose nearly 1,400 members are private and public sector organizations. ANSI is located at 11 West 42nd Street, New York, NY 10036, USA. See also http://web.ansi.org/. ISO (International Organization for Standardization, Organisation Internationale de Normalisation) is an agency of the United Nations whose members are standards organizations of some 100 member countries (one organization from each country). It develops a wide range of standards for industries in all fields. Established in 1947, its mission is “to promote the development of standardization in the world with a view to facilitating the international exchange of goods and 248 5. Statistical Methods services, and to developing cooperation in the spheres of intellectual, scientific, technological and economic activity.” This is the forum where it is agreed, for example, how thick your bank card should be, so every country in the world follows a compatible standard. The ISO is located at 1, rue de Varemb΄e, CH-1211 Geneva 20, Switzerland, http://www.iso.ch/. ITU (International Telecommunication Union) is another United Nations agency developing standards for telecommunications. Its members are mostly companies that make telecommunications equipment, groups interested in standards, and government agencies involved with telecommunications. The ITU is the successor of an organization founded by some 20 European nations in Paris in 1865. IEC (International Electrotechnical Commission) is a nongovernmental international organization that prepares and publishes international standards for all electrical, electronic, and related technologies. The IEC was founded, in 1906, as a result of a resolution passed at the International Electrical Congress held in St. Louis (USA) in 1904. Its membership consists of more than 50 participating countries, including all the world’s major trading nations and a growing number of industrializing countries. The IEC’s mission is to promote, through its members, international cooperation on all questions of electrotechnical standardization and related matters, such as the assessment of conformity to standards, in the fields of electricity, electronics and related technologies. The IEC is located at 3, rue de Varemb΄e, P.O. Box 131, CH-1211, Geneva 20, Switzerland, http://www.iec.ch/. QIC (http://www.qic.org/) was an international trade association, incorporated in 1987, to encourage and promote the widespread use of quarter-inch tape cartridge technology. One hundred companies around the world were Members or Associates of QIC during its active years. The group became inactive in 1998 after more than 15 million QIC-compatible drives had been installed. Most of the 15 million QIC tape drives in use worldwide were installed in business environments. QIC tape automation solutions enabled capacities well into the terabyte range, providing the hardware data compression and read-write features essential to network backup. 5.7 Facsimile Compression Data compression is especially important when images are transmitted over a communications line because the user is often waiting at the receiver, eager to see something quickly. Documents transferred between fax machines are sent as bitmaps, so a standard data compression method was needed when those machines became popular. Several methods were developed and proposed by the ITU-T. The ITU-T is one of four permanent parts of the International Telecommunications Union (ITU), based in Geneva, Switzerland (http://www.itu.ch/). It issues recommendations for standards applying to modems, packet switched interfaces, V.24 connectors, and similar devices. Although it has no power of enforcement, the standards it recommends are generally accepted and adopted by industry. Until March 1993, the ITU-T 5.7 Facsimile Compression 249 CCITT: Can’t Conceive Intelligent Thoughts Today was known as the Consultative Committee for International Telephone and Telegraph (Comit΄e Consultatif International T΄el΄egraphique et T΄el΄ephonique, or CCITT). The first data compression standards developed by the ITU-T were T2 (also known as Group 1) and T3 (Group 2). These are now obsolete and have been replaced by T4 (Group 3) and T6 (Group 4). Group 3 is currently used by all fax machines designed to operate with the Public Switched Telephone Network (PSTN). These are the machines we have at home, and at the time of writing, they operate at maximum speeds of 9,600 baud. Group 4 is used by fax machines designed to operate on a digital network, such as ISDN. They have typical speeds of 64K baud. Both methods can produce compression factors of 10 or better, reducing the transmission time of a typical page to about a minute with the former, and a few seconds with the latter. Some references for facsimile compression are [Anderson et al. 87], [Hunter and Robinson 80], [Marking 90], and [McConnell 92]. 5.7.1 One-Dimensional Coding A fax machine scans a document line by line, converting each line to small black and white dots called pels (from Picture ELement). The horizontal resolution is always 8.05 pels per millimeter (about 205 pels per inch). An 8.5-inch-wide scan line is therefore converted to 1728 pels. The T4 standard, though, recommends to scan only about 8.2 inches, thereby producing 1664 pels per scan line (these numbers, as well as those in the next paragraph, are all to within ±1% accuracy). The vertical resolution is either 3.85 scan lines per millimeter (standard mode) or 7.7 lines/mm (fine mode). Many fax machines have also a very-fine mode, where they scan 15.4 lines/mm. Table 5.29 assumes a 10-inch-high page (254 mm), and shows the total number of pels per page, and typical transmission times for the three modes without compression. The times are long, illustrating how important data compression is in fax transmissions. Scan Pels per Pels per Time Time lines line page (sec.) (min.) 978 1664 1.670M 170 2.82 1956 1664 3.255M 339 5.65 3912 1664 6.510M 678 11.3 Ten inches equal 254 mm. The number of pels is in the millions, and the transmission times, at 9600 baud without compression, are between 3 and 11 minutes, depending on the mode. However, if the page is shorter than 10 inches, or if most of it is white, the compression factor can be 10 or better, resulting in transmission times of between 17 and 68 seconds. Table 5.29: Fax Transmission Times. 250 5. Statistical Methods To derive the Group 3 code, the ITU-T counted all the run lengths of white and black pels in a set of eight “training” documents that they felt represent typical text and images sent by fax, and used the Huffman algorithm to assign a variable-length code to each run length. (The eight documents are described in Table 5.30. They are not shown because they are copyrighted by the ITU-T.) The most common run lengths were found to be 2, 3, and 4 black pixels, so they were assigned the shortest codes (Table 5.31). Next come run lengths of 2–7 white pixels, which were assigned slightly longer codes. Most run lengths were rare and were assigned long, 12-bit codes. Thus, Group 3 uses a combination of RLE and Huffman coding. Image Description 1 Typed business letter (English) 2 Circuit diagram (hand drawn) 3 Printed and typed invoice (French) 4 Densely typed report (French) 5 Printed technical article including figures and equations (French) 6 Graph with printed captions (French) 7 Dense document (Kanji) 8 Handwritten memo with very large white-on-black letters (English) Table 5.30: The Eight CCITT Training Documents.  Exercise 5.18: A run length of 1664 white pels was assigned the short code 011000. Why is this length so common? Since run lengths can be long, the Huffman algorithm was modified. Codes were assigned to run lengths of 1 to 63 pels (they are the termination codes in Table 5.31a) and to run lengths that are multiples of 64 pels (the make-up codes in Table 5.31b). Group 3 is therefore a modified Huffman code (also called MH). The code of a run length is either a single termination code (if the run length is short) or one or more make-up codes, followed by one termination code (if it is long). Here are some examples: 1. A run length of 12 white pels is coded as 001000. 2. A run length of 76 white pels (= 64 + 12) is coded as 11011|001000. 3. A run length of 140 white pels (= 128 + 12) is coded as 10010|001000. 4. A run length of 64 black pels (= 64 + 0) is coded as 0000001111|0000110111. 5. A run length of 2561 black pels (2560 + 1) is coded as 000000011111|010.  Exercise 5.19: There are no runs of length zero. Why then were codes assigned to runs of zero black and zero white pels?  Exercise 5.20: An 8.5-inch-wide scan line results in 1728 pels, so how can there be a run of 2561 consecutive pels? Each scan line is coded separately, and its code is terminated by the special 12-bit EOL code 000000000001. Each line also gets one white pel appended to it on the left when it is scanned. This is done to remove any ambiguity when the line is decoded on 5.7 Facsimile Compression 251 the receiving end. After reading the EOL for the previous line, the receiver assumes that the new line starts with a run of white pels, and it ignores the first of them. Examples: 1. The 14-pel line is coded as the run lengths 1w 3b 2w 2b 7w EOL, which become 000111|10|0111|11|1111|000000000001. The decoder ignores the single white pel at the start. 2. The line is coded as the run lengths 3w 5b 5w 2b EOL, which becomes the binary string 1000|0011|1100|11|000000000001.  Exercise 5.21: The group 3 code for a run length of five black pels (0011) is also the prefix of the codes for run lengths of 61, 62, and 63 white pels. Explain this. The Group 3 code has no error correction, but many errors can be detected. Because of the nature of the Huffman code, even one bad bit in the transmission can cause the receiver to get out of synchronization, and to produce a string of wrong pels. This is why each scan line is encoded separately. If the receiver detects an error, it skips bits, looking for an EOL. This way, one error can cause at most one scan line to be received incorrectly. If the receiver does not see an EOL after a certain number of lines, it assumes a high error rate, and it aborts the process, notifying the transmitter. Since the codes are between 2 and 12 bits long, the receiver detects an error if it cannot decode a valid code after reading 12 bits. Each page of the coded document is preceded by one EOL and is followed by six EOL codes. Because each line is coded separately, this method is a one-dimensional coding scheme. The compression ratio depends on the image. Images with large contiguous black or white areas (text or black and white images) can be highly compressed. Images with many short runs can sometimes produce negative compression. This is especially true in the case of images with shades of gray (such as scanned photographs). Such shades are produced by halftoning, which covers areas with many alternating black and white pels (runs of length one).  Exercise 5.22: What is the compression ratio for runs of length one (i.e., strictly alternating pels)? The T4 standard also allows for fill bits to be inserted between the data bits and the EOL. This is done in cases where a pause is necessary, or where the total number of bits transmitted for a scan line must be a multiple of 8. The fill bits are zeros. Example: The binary string 000111|10|0111|11|1111|000000000001 becomes 000111|10|0111|11|1111|00|0000000001 after two zeros are added as fill bits, bringing the total length of the string to 32 bits (= 8 Χ 4). The decoder sees the two zeros of the fill, followed by the 11 zeros of the EOL, followed by the single 1, so it knows that it has encountered a fill followed by an EOL. See http://www.doclib.org/rfc/rfc804.html for a description of group 3. At the time of writing, the T.4 and T.6 recommendations can also be found at URL ftp://sunsite.doc.ic.ac.uk/ as files 7_3_01.ps.gz and 7_3_02.ps.gz (the precise subdirectory seems to change every few years and it is recommended to locate it with a search engine). 252 5. Statistical Methods (a) White Black White Black Run code- code- Run code- codelength word word length word word 0 00110101 0000110111 32 00011011 000001101010 1 000111 010 33 00010010 000001101011 2 0111 11 34 00010011 000011010010 3 1000 10 35 00010100 000011010011 4 1011 011 36 00010101 000011010100 5 1100 0011 37 00010110 000011010101 6 1110 0010 38 00010111 000011010110 7 1111 00011 39 00101000 000011010111 8 10011 000101 40 00101001 000001101100 9 10100 000100 41 00101010 000001101101 10 00111 0000100 42 00101011 000011011010 11 01000 0000101 43 00101100 000011011011 12 001000 0000111 44 00101101 000001010100 13 000011 00000100 45 00000100 000001010101 14 110100 00000111 46 00000101 000001010110 15 110101 000011000 47 00001010 000001010111 16 101010 0000010111 48 00001011 000001100100 17 101011 0000011000 49 01010010 000001100101 18 0100111 0000001000 50 01010011 000001010010 19 0001100 00001100111 51 01010100 000001010011 20 0001000 00001101000 52 01010101 000000100100 21 0010111 00001101100 53 00100100 000000110111 22 0000011 00000110111 54 00100101 000000111000 23 0000100 00000101000 55 01011000 000000100111 24 0101000 00000010111 56 01011001 000000101000 25 0101011 00000011000 57 01011010 000001011000 26 0010011 000011001010 58 01011011 000001011001 27 0100100 000011001011 59 01001010 000000101011 28 0011000 000011001100 60 01001011 000000101100 29 00000010 000011001101 61 00110010 000001011010 30 00000011 000001101000 62 00110011 000001100110 31 00011010 000001101001 63 00110100 000001100111 (b) White Black White Black Run code- code- Run code- codelength word word length word word 64 11011 0000001111 1344 011011010 0000001010011 128 10010 000011001000 1408 011011011 0000001010100 192 010111 000011001001 1472 010011000 0000001010101 256 0110111 000001011011 1536 010011001 0000001011010 320 00110110 000000110011 1600 010011010 0000001011011 384 00110111 000000110100 1664 011000 0000001100100 448 01100100 000000110101 1728 010011011 0000001100101 512 01100101 0000001101100 1792 00000001000 same as 576 01101000 0000001101101 1856 00000001100 white 640 01100111 0000001001010 1920 00000001101 from this 704 011001100 0000001001011 1984 000000010010 point 768 011001101 0000001001100 2048 000000010011 832 011010010 0000001001101 2112 000000010100 896 011010011 0000001110010 2176 000000010101 960 011010100 0000001110011 2240 000000010110 1024 011010101 0000001110100 2304 000000010111 1088 011010110 0000001110101 2368 000000011100 1152 011010111 0000001110110 2432 000000011101 1216 011011000 0000001110111 2496 000000011110 1280 011011001 0000001010010 2560 000000011111 Table 5.31: Group 3 and 4 Fax Codes: (a) Termination Codes, (b) Make-Up Codes. 5.7 Facsimile Compression 253 5.7.2 Two-Dimensional Coding Two-dimensional coding was developed because one-dimensional coding does not produce good results for images with gray areas. Two-dimensional coding is optional on fax machines that use Group 3 but is the only method used by machines intended to work in a digital network. When a fax machine using Group 3 supports two-dimensional coding as an option, each EOL is followed by one extra bit, to indicate the compression method used for the next scan line. That bit is 1 if the next line is encoded with one-dimensional coding, and 0 if it is encoded with two-dimensional coding. The two-dimensional coding method is also called MMR, for modified modified READ, where READ stands for relative element address designate. The term “modified modified” is used because this is a modification of one-dimensional coding, which itself is a modification of the original Huffman method. The two-dimensional coding method works by comparing the current scan line (called the coding line) to its predecessor (which is called the reference line) and recording the differences between them, the assumption being that two consecutive lines in a document will normally differ by just a few pels. The method assumes that there is an all-white line above the page, which is used as the reference line for the first scan line of the page. After coding the first line, it becomes the reference line, and the second scan line is coded. As in one-dimensional coding, each line is assumed to start with a white pel, which is ignored by the receiver. The two-dimensional coding method is less reliable than one-dimensional coding, since an error in decoding a line will cause errors in decoding all its successors and will propagate through the entire document. This is why the T.4 (Group 3) standard includes a requirement that after a line is encoded with the one-dimensional method, at most K-1 lines will be encoded with the two-dimensional coding method. For standard resolution K = 2, and for fine resolution K = 4. The T.6 standard (Group 4) does not have this requirement, and uses two-dimensional coding exclusively. Scanning the coding line and comparing it to the reference line results in three cases, or modes. The mode is identified by comparing the next run length on the reference line [(b1b2) in Figure 5.33] with the current run length (a0a1) and the next one (a1a2) on the coding line. Each of these three runs can be black or white. The three modes are as follows (see also the flow chart of Figure 5.34): 1. Pass mode. This is the case where (b1b2) is to the left of (a1a2) and b2 is to the left of a1 (Figure 5.33a). This mode does not include the case where b2 is above a1. When this mode is identified, the length of run (b1b2) is coded using the codes of Table 5.32 and is transmitted. Pointer a0 is moved below b2, and the four values b1, b2, a1, and a2 are updated. 2. Vertical mode. (b1b2) overlaps (a1a2) by not more than three pels (Figure 5.33b1, b2). Assuming that consecutive lines do not differ by much, this is the most common case. When this mode is identified, one of seven codes is produced (Table 5.32) and is transmitted. Pointers are updated as in case 1 above. The performance of the twodimensional coding method depends on this case being common. 3. Horizontal mode. (b1b2) overlaps (a1a2) by more than three pels (Figure 5.33c1, c2). When this mode is identified, the lengths of runs (a0a1) and (a1a2) are coded using the codes of Table 5.32 and are transmitted. Pointers are updated as in cases 1 and 2 above. 254 5. Statistical Methods . . . and you thought “impressive” statistics were 36–24–36. —Advertisement, The American Statistician, November 1979 Run length to Abbre- Mode be encoded viation Codeword Pass b1b2 P 0001+coded length of b1b2 Horizontal a0a1, a1a2 H 001+coded length of a0a1 and a1a2 Vertical a1b1 = 0 V(0) 1 a1b1 = -1 VR(1) 011 a1b1 = -2 VR(2) 000011 a1b1 = -3 VR(3) 0000011 a1b1 = +1 VL(1) 010 a1b1 = +2 VL(2) 000010 a1b1 = +3 VL(3) 0000010 Extension 0000001000 Table 5.32: 2D Codes for the Group 4 Method. When scanning starts, pointer a0 is set to an imaginary white pel on the left of the coding line, and a1 is set to point to the first black pel on the coding line. (Recall that a0 corresponds to an imaginary pel, which is why the first run length is |a0a1|-1.) Pointer a2 is set to the first white pel following that. Pointers b1, b2 are set to point to the start of the first and second runs on the reference line, respectively. After identifying the current mode and transmitting codes according to Table 5.32, a0 is updated as shown in the flow chart, and the other four pointers are updated relative to the new a0. The process continues until the end of the coding line is reached. The encoder assumes an extra pel on the right of the line, with a color opposite that of the last pel. The extension code in Table 5.32 is used to abort the encoding process prematurely, before reaching the end of the page. This is necessary if the rest of the page is transmitted in a different code or even in uncompressed form.  Exercise 5.23: Manually figure out the code generated from the two lines below. Ref. line Cod. line Table 5.36 summarizes the codes emitted by the group 4 encoder. Figure 5.35 is a tree with the same codes. Each horizontal branch corresponds to another zero and each vertical branch, to another 1. Teamwork is the ability to work as a group toward a common vision, even if that vision becomes extremely blurry. —Anonymous 5.7 Facsimile Compression 255 (a) Reference line › Coding line › b1 v b2 v a^0 b 1b2 ^ a1 ^ a2 Run length b1b2 coded. New a0 becomes old b2. (b1) Reference line › Coding line › b1 v b2 v ^ a0 ^ a1 ^ a2 a1b1 a1 left of b1 (b2) Reference line › Coding line › b1 v b2 v ^ a0 a1b1  ^ a1 ^ a2 a1 right of b1 Run length a1b1 coded. New a0 becomes old a1. (c1) Reference line › Coding line › b1 v b2 v ^ a0 ^ a1 ^ a2 a0a1  a1a2  (c2) Reference line › Coding line › b1 v b2 v ^ a0 ^ a1 ^ a2  a0a1  a1a2  Run lengths a0a1 (white) and a1a2 (black) coded. New a0 becomes old a2. Notes: 1. a0 is the first pel of a new codeword and can be black or white. 2. a1 is the first pel to the right of a0 with a different color. 3. a2 is the first pel to the right of a1 with a different color. 4. b1 is the first pel on the reference line to the right of a0 with a different color. 5. b2 is the first pel on the reference line to the right of b1 with a different color. Figure 5.33: Five Run-Length Configurations: (a) Pass Mode, (b) Vertical Mode, (c) Horizontal Mode. 256 5. Statistical Methods Start END Reference line := all white Read coding line. Set a0 before 1st pel. b1, and b2 Determine a1, Compute Pass mode code Set a0 under b2 Determine a2 Compute Vertical mode code Set a0 to a1 Compute Horizontal mode code Set a0 to a2 Reference line equal to coding line b2 to the left of a1? No Yes No No Yes Yes No Yes a1b1.3 EOL EOP Figure 5.34: MMR Flow Chart. 5.7 Facsimile Compression 257 0 0 0 0 0 0 0 0 0 0 0 1 1 1 1 1 1 1 1 1 V(0) H P VR(2) VR(3) EOL 0 1 VR(1) VL(1) VL(2) 1 VL(3) 1 2D Extensions 1D Extensions Figure 5.35: Tree of Group 3 Codes. Mode Elements to Be Coded Notation Codeword Pass b1, b2 P 0001 Horizontal a0a1, a1a2 H 001 +M(a0a1) +M(a1a2) a1 just under b1 a1b1 = 0 V(0) 1 a1 to a1b1 = 1 VR(1) 011 the right a1b1 = 2 VR(2) 000011 Vertical of b1 a1b1 = 3 VR(3) 0000011 a1 to a1b1 = 1 VL(1) 010 the left a1b1 = 2 VL(2) 000010 of b1 a1b1 = 3 VL(3) 0000010 2D Extensions 1D Extensions 0000001xxx 000000001xxx EOL 000000000001 1D Coding of Next Line 2D Coding of Next Line EOL+‘1’ EOL+‘0’ Table 5.36: Group 4 Codes. 258 5. Statistical Methods 5.8 PK Font Compression Obviously, these words are hard to read because the individual characters feature different styles and sizes. In a beautifully-typeset document, all the letters of a word and all the words of the document (with perhaps a few exceptions) should have the same size and style; they should belong to the same font. Traditionally, the term “font” refers to a set of characters of type that are of the same size and style, such as Times Roman 12 point. A typeface is a set of fonts of different sizes but in the same style, like Times Roman. A typeface family is a set of typefaces in the same style, such as Times. The size of a font is normally measured in points (more accurately, printer’s points, where 72.27 points equal one inch). The style of a font describes its appearance. Traditional styles are roman, boldface, italic, slanted, typewriter, and sans serif. Old operating systems did not support fonts. Even the DOS operating system, which was widely used on IBM-PC-compatible personal computers from 1980 to 1995, did not support fonts and was based on ASCII codes instead. It was not until the mid 1980s that digital fonts became a part of many operating systems. In 1984, Adobe launched the PostScript language, which supported two types of digital fonts. In the same year, Apple released the first models of the Macintosh computer, whose operating system used fonts (Figure 5.37 illustrates some of the early fonts included in the Macintosh OS 6). In 1985, Apple announced the LaserWriter, one of the first laser printers intended for the mass market. These developments paved the way for future operating systems to support digital fonts, thus enabling users to create, print, and publish beautiful, professionally-looking documents. Figure 5.37: Old Macintosh OS 6 Fonts. A digital font is a set of symbols or characters that can be stored in the computer’s memory or on an I/O device such as a disk. The symbols of the font are either displayed or printed. Each symbol consists of a glyph (the actual shape of the symbol) and bookkeeping information, such as the height, depth, and width of the symbol. Modern fonts are referred to as outline fonts. A symbol in such a font is fully defined by a small set of control points that are connected by curves to form the outline of the symbol. The symbol . in Figure 5.38 is an example. It is easy to see how this symbol is fully specified by about 30 control points that are connected with a few smooth curves. An outline font is easy to store in a computer file simply by storing the coordinates of the control points. It is also easy to change the size of the font by scaling the coordinates of the points. In contrast, early digital fonts were of the bitmap variety. The font designer would draw the glyph of each symbol on a grid of small squares (pixels) and then select the 5.8 PK Font Compression 259 xxxxxxxxxxxxxxxxxxxx xxxxxxxxxxxxxxxxxxxx xxxxxxxxxxxxxxxxxxxx xxxxxxxxxxxxxxxxxxxx xx................xx xx................xx xx................xx .................... .................... ..xx............xx.. ..xx............xx.. ..xx............xx.. ..xxxxxxxxxxxxxxxx.. ..xxxxxxxxxxxxxxxx.. ..xxxxxxxxxxxxxxxx.. ..xxxxxxxxxxxxxxxx.. ..xxxxxxxxxxxxxxxx.. ..xx............xx.. ..xx............xx.. ..xx............xx.. .................... .................... .................... xx................xx xx................xx xx................xx xxxxxxxxxxxxxxxxxxxx xxxxxxxxxxxxxxxxxxxx xxxxxxxxxxxxxxxxxxxx xxxxxxxxxxxxxxxxxxxx xxxxxxxxxxxxxxxxxx ..xxxxxxxxxxxxxxxx ...xxxx.........xx ...xxx..........xx ...xxx...........x ...xxx...........x ...xxx............ ...xxx............ ...xxx.........x.. ...xxx.........x.. ...xxx.........x.. ...xxx........xx.. ...xxxxxxxxxxxxx.. ...xxxxxxxxxxxxx.. ...xxx........xx.. ...xxx........xx.. ...xxx.........x.. ...xxx.........x.. ...xxx............ ...xxx............ ...xxx............ ...xxx............ ...xxx............ ...xxx............ ...xxx............ ..xxxxx........... xxxxxxxxx......... .........xx......... ......xxxxxxxx...... .....xx.....xxx..... ....xxx.......xx.... ...xxx........xxx... ..xxxx.........xx... ..xxx...........xx.. .xxxx...........xxx. .xxx............xxx. .xxx............xxx. .xxx............xxx. .xxx............xxx. .xxx............xxx. .xxx............xxx. ..xxx..........xxx.. ..xxx..........xxx.. ...xxx........xxxx.. x...xx.......xxx...x .x....xx....x.....x. .x.....xx...x.....x. .xx....xx...x.....x. .xxxxxxxx...xxxxxxx. .xxxxxxxx...xxxxxxx. xxxxxxxxx..xxxxxxxxx ..xxxxx......xxxxx.. ...xxx.......xxxx... ...xxx........xxx... ...xxx........xxx... ...xxx........xxx... ...xxx........xxx... ...xxx........xxx... ...xxx........xxx... ..,xxx........xxx... ...xxxx.......xxx... ...xxxxx.....xxxx... ....xxxxx...xxxxx... .....xxxxxxxx.xxx... ......xxxxxx..xxx... ..............xxx... ..............xxx... ..............xxx... ..............xxx... ..............xxx... ..............xxx... ..............xxx... .............xxxxx.. ...........xxxxxxxxx .................xxx ..................xx ..................xx ...................x ...................x Figure 5.38: Outline and Bitmap Fonts. best pixels for the symbol. Figure 5.38 shows four examples of bitmap symbols where the white pixels are represented by dots and the black pixels are shown as an Χ. In the computer’s memory, a bitmap symbol is stored as a map of bits, zeros for white pixels and 1’s for black pixels. When such a font is written in raw format on a file, the file tends to be large and should be compressed. Also, a high-quality bitmap font has to be designed from scratch for each font size. Simply scaling a bitmap results in badly-looking characters. The TEX project, started by Donald Knuth in 1975, is an early example of the use of bitmap fonts. The goal was to develop software for typesetting beautiful documents, especially documents with a lot of mathematics. Such software requires a set of fonts, even if the operating system does not support fonts. Thus, METAFONT, an application to generate fonts, was developed in parallel with the TEX application itself. METAFONT was then used to create a set of fonts that is referred to as computer modern (CM). The CM set of fonts consists of 75 fonts that are described in volume E of [Knuth 86] as well as the “line,” “circle,” and symbol fonts associated with LaTeX. The output of METAFONT is called a generic font (GF), to indicate that this file format does not follow the conventions of any font foundry. However, it is easy to convert GF font files to the special format required by most digital phototypesetting equipment. In 1985, Tomas Rokicki, a student of Knuth’s, developed the PK (for PacKed) font format. The idea was to develop a simple, fast compression method such that the fonts would be saved in small files and even the slow computers available at that time would be able to decompress a font, or any part of it, quickly. The main references for the format and organization of PK fonts are [Rokicki 95], [Rokicki 90], and [Haralambous 07]. The method selected for PK compression was based on run-length encoding (RLE) and on the special features of font bitmaps. The method is not very efficient, but it is described here because it offers an original approach to RLE, an approach that does not use Huffman codes and makes minimal use of variable-length codes (the packed numbers, described later, are the only variable-length code used by PK). A quick glance at the four bitmaps of Figure 5.38 shows two important features (1) they are narrow, resulting in mostly short runs of black and white pixels and (2) they tend to have consecutive identical rows of pixels. This section discusses the compression of PK fonts, but a full understanding of this compression method requires a little knowledge of the organization of these fonts and of the way TEX employs font information. When a font of type is designed, the designer determines, for each symbol in the font, its glyph and its dimensions. Symbols in a font are two-dimensional, but each has 260 5. Statistical Methods three dimensions, height, depth, and width. It is best to think of the symbol as if it is surrounded by a box (or a bitmap) as illustrated by the leftmost item of Figure 5.39. There is a baseline that separates the height and depth of the box, and there is a reference point at the left end of the baseline. h w d baseline Figure 5.39: Three-Dimensional Character Boxes. TEX uses the three dimensions of each font symbol but is not concerned with the actual shape of the symbol. As a result, symbols may stick out of their bitmap boxes, as illustrated by the italic d in the figure. TEX reads the input text and strings boxes horizontally to construct a line of text. The boxes butt together (in Figure 5.39 there are small spaces between boxes for better readability), so the font designer leaves spaces to the left and right of each symbol, as illustrated in the figure. (In those rare cases where a symbol, such as an m-dash, should touch their neighbors, the bitmap box is as wide as the symbol and no spaces are left.) When the current line of text is ready, TEX starts on the next line. When that line is also ready, it is placed under the current line (and it then becomes the new current line) such that the baselines of the two lines are separated by a user-controlled parameter. If this causes the lines to overlap (because the top line has a symbol with a large depth and the bottom line has a symbol with a large height), TEX leaves some space (Figure 5.40) between the bottom of the upper line (the symbol with the largest depth) and the top of the lower line (the symbol with the largest height). This space can also be controlled by the user, but the font designer does not have to worry about it. Thus, there are spaces to the left and right of font symbols in their bitmap boxes but no spaces above and below them. Figure 5.40: Two Lines of Text. PK compression is based on run-length encoding. The method encodes the runs of black and white pixels in the bitmap of a symbol. In order to decrease the run 5.8 PK Font Compression 261 lengths and thereby increase compression performance, the encoder first determines the bounding box of a bitmap, as illustrated by the 16 Χ 20 bitmap of Figure 5.40. The bounding box of a glyph bitmap is the smallest rectangle that encompasses all the black pixels of the glyph. It is easy to see that our bounding box is 15 pixels wide and 16 pixels tall. It starts on column 3 from the left and 13 of the 16 pixel rows are above the baseline. Once the bounding box is decompressed, the decoder needs the following dimensions to create the complete bitmap: The horizontal escapement (the width of the bitmap, 20), the width of the bounding box (15), the vertical escapement (the height of the bitmap, 16), the height of the bounding box (also 16), the horizontal offset (the distance between the reference point and leftmost column of the bounding box, 2), and the vertical offset (the distance from the reference point to the top of the bounding box, 13). Three of these dimensions are horizontal and three are vertical, but often only two vertical dimensions are needed, because the bitmap and bounding box tend to have the same height. We are now ready to look at the details of the compression. A PK font file is organized in records, one for each symbol. A record starts with a character preamble that contains (in addition to other information) the five dimensions mentioned above and one bit that specifies the top-left pixel of the bounding box (1 for a black pixel and 0 for a white one). The character preamble is followed by the encoded run lengths of the symbol’s pixels and by the preamble of the next character. Because of the use of nybbles to encode run lengths, the preamble may start in the middle of a byte. There are also a preamble and postamble for the entire file. The first step of the encoder is to locate groups of repeating (i.e., identical consecutive) pixel rows in the bounding box. When a group of n such rows is located, n - 1 of them are removed from the bounding box, and a repeat count of [n - 1] is suitably encoded somewhere between runs in the remaining row of the group. Figure 5.41 illustrates this process. Part (a) of the figure shows a 5Χ15 bitmap (note that this is not a bounding box) where the three middle rows repeat. In part (b), two of these rows have been deleted and part (c) shows the valid points between runs where the repeat count [2] may be inserted. 000001111000000 011100111101100 011100111101100 011100111101100 000111111000011 000001111000000 011100111101100 000111111000011 000001111000000 011100111101100 ^ ^ ^ ^^ ^ 000111111000011 000001111000000 000000000000000 000000000000000 000000000000000 000000000000000 110000000000000 (a) (b) (c) (d) Figure 5.41: Repeating Rows in a Bitmap. Thus, the run lengths of this example become the following sequence 5, 4, 7, [2], 3, 2, 4, 1, 2, 5, 6, 4, and 2. The repeat count was placed at the first valid point, but could have been placed at any of the other five points indicated in the figure. The repeat count for a group of identical rows is placed in the remaining row between runs of pixels, but this rule fails in bitmaps such as the one of Figure 5.41d, where the repeating rows are uniform and are preceded by a run of the same color pixels. In such a case, the encoder 262 5. Statistical Methods does not delete the repeating rows and ends up encoding a long run (66 white pixels in the figure), which reduces the overall compression somewhat. It is now clear that we have to encode two types of data, run lengths of pixels and repeat counts. Since a typical bounding box is small, we expect most run lengths to be short and most repeat counts to be small. Thus, the principle of PK compression is to pack, whenever possible, two items in the two nybbles of a byte. Often, the first nybble is a flag and the second nybble is a data item. A nybble is four bits long and can have 16 values. We expect to have more run lengths than repeat counts, so we reserve nybble values 14 and 15 to indicate repeat counts. A nybble of 15 indicates a repeat count of 1 (the most common) and a nybble of 14 will be followed by a repeat count stored as a packed number (a term to be discussed shortly). The remaining nybble values 0 through 13 indicate run lengths. Value 0 indicates a long run, occupying three or more nybbles. The length of the run is stored in as many nybbles as needed, encoded as a packed number. Values 1 through 13 indicate the short and medium run lengths. In order to use these 13 values to maximum effect, the encoder counts the run lengths of the bounding box of the current character, and uses their values, in a fast, one-pass algorithm, to compute a variable dyn. This variable is computed for each character of the font and is stored in the character’s preamble. Variable dyn is used as follows: run lengths 1 through dyn are considered short and are stored in one nybble each. The medium run lengths are stored in two nybbles, the first of which, to be denoted by a, is between dyn + 1 and 13 and the second one, b, can be any 4-bit value. The run length represented by a and b is defined as 16(a-dyn-1)+b+dyn+1. The shortest run length is therefore dyn+1 (for a = dyn+1 and b = 0) and the longest is 16(13 - dyn) + dyn (for a = 13 and b = 15). This convention is best illustrated by examples. We first select dyn = 4. In this case, run lengths 1 through 4 are the short ones and are represented by one nybble each, thus 0001, 0010, 0011, and 0100. The medium run lengths are 5 (for a = 4+1 and b = 0) through 16(13-4)+4 = 148 (for a = 13 and b = 15). Runs of more than 148 pixels are considered long and will start with a zero nybble. We now select dyn = 12. In this case, run lengths 1 through 12 are short and are represented by one nybble each. The medium run lengths are 13 (for a = 12 + 1 and b = 0) through 16(13 - 12) + 12 = 28 (for a = 13 and b = 15). Runs of more than 28 pixels are considered long. It is clear that the value of dyn is critical. Small dyn values allow for a large range of medium run lengths, while large values of dyn should be used in cases where most run lengths are short. Next, we discuss the format of a packed number. Given an integer i, the idea is to create its hexadecimal representation (let’s say it occupies n nybbles), remove any leading zero nybbles and prepend n-1 zero nybbles. Thus, i values 1 through 15 occupy one nybble (nothing is prepended). Values 16 . i . 255 are two nybbles long, so one zero nybble should be prepended, for a total of three nybbles. Values 256 . i . 4,095 occupy three nybbles, so two nybbles should be prepended, for a total of five nybbles. This variable-length format is used to encode repeat counts (indicated by a nybble flag of 14) and the long run lengths (indicated by a nybble flag of 0). However, the long run lengths are always greater than 16(13 - dyn) + dyn. The shortest of the long run 5.8 PK Font Compression 263 lengths is therefore s def = 16(13-dyn)+dyn+1, which is why it makes sense to subtract s from such a run length before it is encoded as a packed number. Recall that the long run lengths are indicated by a nybble flag of 0, and a packed integer in the interval [16, 255] is also preceded by a single zero nybble. Thus, it makes sense to add 16 to the long run lengths after s is subtracted. Example. If dyn = 4, the longest medium run length is 148, so s = 149. Thus, a run length of 200 is long and is encoded by computing 200 - 149 + 16 = 67 = 4316 and constructing the 3-nybble packed number 04316. The algorithm for determining the optimal value of dyn (between 1 and 13) can now be described. This algorithm is executed after the sequence of run lengths has been determined. We start with a naive version of the algorithm. For each value of dyn between 1 and 13, it is easy to compute the ranges of the short, medium, and long runs. Each run in the sequence is examined, its type (short, medium, or long) is easily determined, and its length after being encoded is computed (short runs are one nybble long, medium runs are two nybbles, and long runs occupy three or more nybbles and their lengths take a bit more work to determine). The 13 total lengths for the values of dyn are saved and their minimum is then found. A better version of this algorithm exploits the fact that as dyn is increased, the range of short run lengths increases while the range of medium run lengths decreases (for dyn = 4 this range is [5, 148] but for dyn = 12 it is only [13, 28]). Thus, a run of pixels that was medium for a certain value of dyn may remain medium but may also become short or long when dyn is incremented by 1. This version of the algorithm starts by setting dyn to zero. Thus, initially there are no short runs. The algorithm goes over all the run lengths. It determines the type and computes the encoded length of each run. The lengths are added into variable total. The algorithm then increments dyn by 1 and goes over all the runs again. If a run was short in the previous iteration, it will remain short. If it was medium and has now become short, it decreases the value of total by 1. If it was medium and has now become long, it increases the value of total by 1 (or in rare cases, by 2). After each iteration, the new value of total is saved in an array. After 13 iterations, the smallest total is located in the array and is used to determine the optimal value of dyn. The PK compression method described so far is simple, but not very efficient. Given a bitmap with many short runs (such as a gray character, where black and white pixels alternate), this RLE-based method may produce large expansion. In such cases, the run length encoding described here is abandoned and the bitmap is simply written on the font file in raw format (one bit per pixel). We end with a summarizing example. The middle bitmap of Figure 5.38 is the Greek letter . (Xi). Its size is 20Χ29 pixels and it has several groups of repeating rows, although only five groups are not uniform and can employ repeat counts. Counting the runs of pixels produces the sequence 82, [2], 16, 2, 42, [2], 2, 12, 2, 4, [3], 16, 4, [2], 2, 12, 2, 62, [2], 2, 16, and 82. The optimal value of dyn turns out to be 8, but in order to make this example useful and have short, medium, and long runs, we assume that dyn is set to 12. Thus, the short runs are 1 to 12 pixels, the medium runs are 13 to 28 pixels, and the long runs are longer than 28 pixels. The first run is 82; a long run. Subtracting s = 29 and adding 16 yields 69 = 4516. This is encoded as the three nybbles 045. The repeat count of 2 is encoded as the nybble 264 5. Statistical Methods 14 = E16, followed by the packed number representation of 2, thus E2. The medium run of 16 is encoded in two nybbles ab, but since a is between dyn+1 and 13, its value must be 13 = D16, implying that b = 4 and the run of 16 is encoded as D4. The next run, 2, is short and is encoded as the single nybble 2. The run 42 is long. We first compute 42-29+16 = 29 = 1D16 and then encode 01D. The next two runs [2] and 2 are encoded as E22. So far we have 045E2D4201DE22. The next run, 12, is short and is encoded as the single nybble C. The remaining runs do not provide any new examples and should be encoded by the reader as an exercise. 5.9 Arithmetic Coding The Huffman method is simple, efficient, and produces the best codes for the individual data symbols. However, Section 5.2 shows that the only case where it produces ideal variable-length codes (codes whose average size equals the entropy) is when the symbols have probabilities of occurrence that are negative powers of 2 (i.e., numbers such as 1/2, 1/4, or 1/8). This is because the Huffman method assigns a code with an integral number of bits to each symbol in the alphabet. Information theory shows that a symbol with probability 0.4 should ideally be assigned a 1.32-bit code, since -log2 0.4 . 1.32. The Huffman method, however, normally assigns such a symbol a code of 1 or 2 bits. Arithmetic coding overcomes the problem of assigning integer codes to the individual symbols by assigning one (normally long) code to the entire input file. The method starts with a certain interval, it reads the input file symbol by symbol, and it uses the probability of each symbol to narrow the interval. Specifying a narrower interval requires more bits, so the number constructed by the algorithm grows continuously. To achieve compression, the algorithm is designed such that a high-probability symbol narrows the interval less than a low-probability symbol, with the result that high-probability symbols contribute fewer bits to the output. An interval can be specified by its lower and upper limits or by one limit and width. We use the latter method to illustrate how an interval’s specification becomes longer as the interval narrows. The interval [0, 1] can be specified by the two 1-bit numbers 0 and 1. The interval [0.1, 0.512] can be specified by the longer numbers 0.1 and 0.412. The very narrow interval [0.12575, 0.1257586] is specified by the long numbers 0.12575 and 0.0000086. The output of arithmetic coding is interpreted as a number in the range [0, 1). [The notation [a, b) means the range of real numbers from a to b, including a but not including b. The range is “closed” at a and “open” at b.] Thus the code 9746509 is be interpreted as 0.9746509, although the 0. part is not included in the output file. Before we plunge into the details, here is a bit of history. The principle of arithmetic coding was first proposed by Peter Elias in the early 1960s and was first described in [Abramson 63]. Early practical implementations of this method were developed by [Rissanen 76], [Pasco 76], and [Rubin 79]. [Moffat et al. 98] and [Witten et al. 87] should especially be mentioned. They discuss both the principles and details of practical arithmetic coding and show examples. The first step is to calculate, or at least to estimate, the frequencies of occurrence of each symbol. For best results, the exact frequencies are calculated by reading the 5.9 Arithmetic Coding 265 entire input file in the first pass of a two-pass compression job. However, if the program can get good estimates of the frequencies from a different source, the first pass may be omitted. The first example involves the three symbols a1, a2, and a3, with probabilities P1 = 0.4, P2 = 0.5, and P3 = 0.1, respectively. The interval [0, 1) is divided among the three symbols by assigning each a subinterval proportional in size to its probability. The order of the subintervals is immaterial. In our example, the three symbols are assigned the subintervals [0, 0.4), [0.4, 0.9), and [0.9, 1.0). To encode the string “a2a2a2a3”, we start with the interval [0, 1). The first symbol a2 reduces this interval to the subinterval from its 40% point to its 90% point. The result is [0.4, 0.9). The second a2 reduces [0.4, 0.9) in the same way (see note below) to [0.6, 0.85), the third a2 reduces this to [0.7, 0.825), and the a3 reduces this to the stretch from the 90% point of [0.7, 0.825) to its 100% point, producing [0.8125, 0.8250). The final code our method produces can be any number in this final range. (Note: The subinterval [0.6, 0.85) is obtained from the interval [0.4, 0.9) by 0.4 + (0.9 - 0.4) Χ 0.4 = 0.6 and 0.4 + (0.9 - 0.4) Χ 0.9 = 0.85.) With this example in mind, it should be easy to understand the following rules, which summarize the main steps of arithmetic coding: 1. Start by defining the “current interval” as [0, 1). 2. Repeat the following two steps for each symbol s in the input stream: 2.1. Divide the current interval into subintervals whose sizes are proportional to the symbols’ probabilities. 2.2. Select the subinterval for s and define it as the new current interval. 3. When the entire input stream has been processed in this way, the output should be any number that uniquely identifies the current interval (i.e., any number inside the current interval). For each symbol processed, the current interval gets smaller, so it takes more bits to express it, but the point is that the final output is a single number and does not consist of codes for the individual symbols. The average code size can be obtained by dividing the size of the output (in bits) by the size of the input (in symbols). Notice also that the probabilities used in step 2.1 may change all the time, since they may be supplied by an adaptive probability model (Section 5.10). A theory has only the alternative of being right or wrong. A model has a third possibility: it may be right, but irrelevant. —Manfred Eigen, The Physicist’s Conception of Nature The next example is a little more involved. We show the compression steps for the short string SWISSMISS. Table 5.42 shows the information prepared in the first step (the statistical model of the data). The five symbols appearing in the input may be arranged in any order. For each symbol, its frequency is first counted, followed by its probability of occurrence (the frequency divided by the string size, 10). The range [0, 1) is then divided among the symbols, in any order, with each symbol getting a chunk, or a subrange, equal in size to its probability. Thus S gets the subrange [0.5, 1.0) (of 266 5. Statistical Methods size 0.5), whereas the subrange of I is of size 0.2 [0.2, 0.4). The cumulative frequencies column is used by the decoding algorithm on page 271. Char Freq Prob. Range CumFreq Total CumFreq= 10 S 5 5/10 = 0.5 [0.5, 1.0) 5 W 1 1/10 = 0.1 [0.4, 0.5) 4 I 2 2/10 = 0.2 [0.2, 0.4) 2 M 1 1/10 = 0.1 [0.1, 0.2) 1  1 1/10 = 0.1 [0.0, 0.1) 0 Table 5.42: Frequencies and Probabilities of Five Symbols. The symbols and frequencies in Table 5.42 are written on the output stream before any of the bits of the compressed code. This table will be the first thing input by the decoder. The encoding process starts by defining two variables, Low and High, and setting them to 0 and 1, respectively. They define an interval [Low, High). As symbols are being input and processed, the values of Low and High are moved closer together, to narrow the interval. After processing the first symbol S, Low and High are updated to 0.5 and 1, respectively. The resulting code for the entire input stream will be a number in this range (0.5 . Code < 1.0). The rest of the input stream will determine precisely where, in the interval [0.5, 1), the final code will lie. A good way to understand the process is to imagine that the new interval [0.5, 1) is divided among the five symbols of our alphabet using the same proportions as for the original interval [0, 1). The result is the five subintervals [0.5, 0.55), [0.55, 0.60), [0.60, 0.70), [0.70, 0.75), and [0.75, 1.0). When the next symbol W is input, the third of those subintervals is selected and is again divided into five subsubintervals. As more symbols are being input and processed, Low and High are being updated according to NewHigh:=OldLow+Range*HighRange(X); NewLow:=OldLow+Range*LowRange(X); where Range=OldHigh-OldLow and LowRange(X), HighRange(X) indicate the low and high limits of the range of symbol X, respectively. In the example above, the second input symbol is W, so we update Low := 0.5 + (1.0 - 0.5) Χ 0.4 = 0.70, High := 0.5 + (1.0-0.5)Χ0.5 = 0.75. The new interval [0.70, 0.75) covers the stretch [40%, 50%) of the subrange of S. Table 5.43 shows all the steps involved in coding the string SWISSMISS (the first three steps are illustrated graphically in Figure 5.56a). The final code is the final value of Low, 0.71753375, of which only the eight digits 71753375 need be written on the output stream (but see later for a modification of this statement). The decoder works in reverse. It starts by inputting the symbols and their ranges, and reconstructing Table 5.42. It then inputs the rest of the code. The first digit is 7, so the decoder immediately knows that the entire code is a number of the form 0.7 . . .. 5.9 Arithmetic Coding 267 Char. The calculation of low and high S L 0.0 + (1.0 - 0.0) Χ 0.5= 0.5 H 0.0 + (1.0 - 0.0) Χ 1.0= 1.0 W L 0.5 + (1.0 - 0.5) Χ 0.4= 0.70 H 0.5 + (1.0 - 0.5) Χ 0.5= 0.75 I L 0.7 + (0.75 - 0.70) Χ 0.2= 0.71 H 0.7 + (0.75 - 0.70) Χ 0.4= 0.72 S L 0.71 + (0.72 - 0.71) Χ 0.5= 0.715 H 0.71 + (0.72 - 0.71) Χ 1.0= 0.72 S L 0.715 + (0.72 - 0.715) Χ 0.5= 0.7175 H 0.715 + (0.72 - 0.715) Χ 1.0= 0.72  L 0.7175 + (0.72 - 0.7175) Χ 0.0= 0.7175 H 0.7175 + (0.72 - 0.7175) Χ 0.1= 0.71775 M L 0.7175 + (0.71775 - 0.7175) Χ 0.1= 0.717525 H 0.7175 + (0.71775 - 0.7175) Χ 0.2= 0.717550 I L 0.717525 + (0.71755 - 0.717525) Χ 0.2= 0.717530 H 0.717525 + (0.71755 - 0.717525) Χ 0.4= 0.717535 S L 0.717530 + (0.717535 - 0.717530) Χ 0.5= 0.7175325 H 0.717530 + (0.717535 - 0.717530) Χ 1.0= 0.717535 S L 0.7175325 + (0.717535 - 0.7175325) Χ 0.5= 0.71753375 H 0.7175325 + (0.717535 - 0.7175325) Χ 1.0= 0.717535 Table 5.43: The Process of Arithmetic Encoding. This number is inside the subrange [0.5, 1) of S, so the first symbol is S. The decoder then eliminates the effect of symbol S from the code by subtracting the lower limit 0.5 of S and dividing by the width of the subrange of S (0.5). The result is 0.4350675, which tells the decoder that the next symbol is W (since the subrange of W is [0.4, 0.5)). To eliminate the effect of symbol X from the code, the decoder performs the operation Code:=(Code-LowRange(X))/Range, where Range is the width of the subrange of X. Table 5.45 summarizes the steps for decoding our example string (notice that it has two rows per symbol). The next example is of three symbols with probabilities as shown in Table 5.46a. Notice that the probabilities are very different. One is large (97.5%) and the others much smaller. This is a case of skewed probabilities. Encoding the string a2a2a1a3a3 produces the strange numbers (accurate to 16 digits) in Table 5.47, where the two rows for each symbol correspond to the Low and High values, respectively. Figure 5.44 lists the Mathematica code that computed the table. At first glance, it seems that the resulting code is longer than the original string, but Section 5.9.3 shows how to figure out the true compression achieved by arithmetic coding. Decoding this string is shown in Table 5.48 and involves a special problem. After eliminating the effect of a1, on line 3, the result is 0. Earlier, we implicitly assumed 268 5. Statistical Methods that this means the end of the decoding process, but now we know that there are two more occurrences of a3 that should be decoded. These are shown on lines 4, 5 of the table. This problem always occurs when the last symbol in the input stream is the one whose subrange starts at zero. In order to distinguish between such a symbol and the end of the input stream, we need to define an additional symbol, the end-of-input (or end-of-file, eof). This symbol should be added, with a small probability, to the frequency table (see Table 5.46b), and it should be encoded at the end of the input stream. lowRange={0.998162,0.023162,0.}; highRange={1.,0.998162,0.023162}; low=0.; high=1.; enc[i_]:=Module[{nlow,nhigh,range}, range=high-low; nhigh=low+range highRange[[i]]; nlow=low+range lowRange[[i]]; low=nlow; high=nhigh; Print["r=",N[range,25]," l=",N[low,17]," h=",N[high,17]]] enc[2] enc[2] enc[1] enc[3] enc[3] Figure 5.44: Mathematica Code for Table 5.47. Tables 5.49 and 5.50 show how the string a3a3a3a3eof is encoded into the number 0.0000002878086184764172, and then decoded properly. Without the eof symbol, a string of all a3s would have been encoded into a 0. Notice how the low value is 0 until the eof is input and processed, and how the high value quickly approaches 0. Now is the time to mention that the final code does not have to be the final low value but can be any number between the final low and high values. In the example of a3a3a3a3eof, the final code can be the much shorter number 0.0000002878086 (or 0.0000002878087 or even 0.0000002878088).  Exercise 5.24: Encode the string a2a2a2a2 and summarize the results in a table similar to Table 5.49. How do the results differ from those of the string a3a3a3a3? If the size of the input stream is known, it is possible to do without an eof symbol. The encoder can start by writing this size (unencoded) on the output stream. The decoder reads the size, starts decoding, and stops when the decoded stream reaches this size. If the decoder reads the compressed stream byte by byte, the encoder may have to add some zeros at the end, to make sure the compressed stream can be read in groups of 8 bits. 5.9 Arithmetic Coding 269 Char. Code-low Range S 0.71753375 - 0.5= 0.21753375/0.5 = 0.4350675 W 0.4350675 - 0.4 =0.0350675 /0.1 = 0.350675 I 0.350675 - 0.2 =0.150675 /0.2 = 0.753375 S 0.753375 - 0.5 =0.253375 /0.5 = 0.50675 S 0.50675 - 0.5 =0.00675 /0.5 = 0.0135  0.0135 -0 =0.0135 /0.1 = 0.135 M 0.135 - 0.1 =0.035 /0.1 = 0.35 I 0.35 - 0.2 =0.15 /0.2 = 0.75 S 0.75 - 0.5 =0.25 /0.5 = 0.5 S 0.5 - 0.5 =0 /0.5 = 0 Table 5.45: The Process of Arithmetic Decoding. Char Prob. Range a1 0.001838 [0.998162, 1.0) a2 0.975 [0.023162, 0.998162) a3 0.023162 [0.0, 0.023162) (a) Char Prob. Range eof 0.000001 [0.999999, 1.0) a1 0.001837 [0.998162, 0.999999) a2 0.975 [0.023162, 0.998162) a3 0.023162 [0.0, 0.023162) (b) Table 5.46: (Skewed) Probabilities of Three Symbols. a2 0.0 + (1.0 - 0.0) Χ 0.023162= 0.023162 0.0 + (1.0 - 0.0) Χ 0.998162= 0.998162 a2 0.023162 + .975 Χ 0.023162= 0.04574495 0.023162 + .975 Χ 0.998162= 0.99636995 a1 0.04574495 + 0.950625 Χ 0.998162= 0.99462270125 0.04574495 + 0.950625 Χ 1.0= 0.99636995 a3 0.99462270125 + 0.00174724875 Χ 0.0= 0.99462270125 0.99462270125 + 0.00174724875 Χ 0.023162= 0.994663171025547 a3 0.99462270125 + 0.00004046977554749998Χ 0.0= 0.99462270125 0.99462270125 + 0.00004046977554749998Χ 0.023162= 0.994623638610941 Table 5.47: Encoding the String a2a2a1a3a3. 270 5. Statistical Methods Char. Code-low Range a2 0.99462270125 - 0.023162= 0.97146170125/0.975 = 0.99636995 a2 0.99636995 - 0.023162 = 0.97320795 /0.975 = 0.998162 a1 0.998162 - 0.998162 = 0.0 /0.00138 = 0.0 a3 0.0 - 0.0 =0.0 /0.023162= 0.0 a3 0.0 - 0.0 =0.0 /0.023162= 0.0 Table 5.48: Decoding the String a2a2a1a3a3. a3 0.0 + (1.0 - 0.0) Χ 0.0= 0.0 0.0 + (1.0 - 0.0) Χ 0.023162= 0.023162 a3 0.0 + .023162 Χ 0.0= 0.0 0.0 + .023162 Χ 0.023162= 0.000536478244 a3 0.0 + 0.000536478244 Χ 0.0= 0.0 0.0 + 0.000536478244 Χ 0.023162= 0.000012425909087528 a3 0.0 + 0.000012425909087528Χ 0.0= 0.0 0.0 + 0.000012425909087528Χ 0.023162= 0.0000002878089062853235 eof 0.0 + 0.0000002878089062853235Χ 0.999999= 0.0000002878086184764172 0.0 + 0.0000002878089062853235Χ 1.0= 0.0000002878089062853235 Table 5.49: Encoding the String a3a3a3a3eof. Char. Code-low Range a3 0.0000002878086184764172-0 =0.0000002878086184764172 /0.023162=0.00001242589666161891247 a3 0.00001242589666161891247-0=0.00001242589666161891247/0.023162=0.000536477707521756 a3 0.000536477707521756-0 =0.000536477707521756 /0.023162=0.023161976838 a3 0.023161976838-0.0 =0.023161976838 /0.023162=0.999999 eof 0.999999-0.999999 =0.0 /0.000001=0.0 Table 5.50: Decoding the String a3a3a3a3eof. 5.9.1 Implementation Details The encoding process described earlier is not practical, since it assumes that numbers of unlimited precision can be stored in Low and High. The decoding process described on page 267 (“The decoder then eliminates the effect of the S from the code by subtracting. . . and dividing . . . ”) is simple in principle but also impractical. The code, which is a single number, is normally long and may also be very long. A 1 Mbyte file may be encoded into, say, a 500 Kbyte file that consists of a single number. Dividing a 500 Kbyte number is complex and slow. Any practical implementation of arithmetic coding should use just integers (because floating-point arithmetic is slow and precision is lost), and they should not be very long 5.9 Arithmetic Coding 271 (preferably just single precision). We describe such an implementation here, using two integer variables Low and High. In our example they are four decimal digits long, but in practice they might be 16 or 32 bits long. These variables hold the low and high limits of the current subinterval, but we don’t let them grow too much. A glance at Table 5.43 shows that once the leftmost digits of Low and High become identical, they never change. We therefore shift such digits out of the two variables and write one digit on the output stream. This way, the two variables don’t have to hold the entire code, just the most-recent part of it. As digits are shifted out of the two variables, a zero is shifted into the right end of Low and a 9 into the right end of High. A good way to understand this is to think of each of the two variables as the left end of an infinitely long number. Low contains xxxx00 . . ., and High= yyyy99 . . . . One problem is that High should be initialized to 1, but the contents of Low and High should be interpreted as fractions less than 1. The solution is to initialize High to 9999. . . , since the infinite fraction 0.999 . . . equals 1. (This is easy to prove. If 0.999 . . . < 1, then their average a = (1+0.999 . . .)/2 would be a number between 0.999 . . . and 1, but there is no way to write a. It is impossible to give it more digits than to 0.999 . . ., since the latter already has an infinite number of digits. It is impossible to make the digits any bigger, since they are already 9’s. This is why the infinite fraction 0.999 . . . must equal 1.)  Exercise 5.25: Write the number 0.5 in binary. Table 5.51 describes the encoding process of the string SWISSMISS. Column 1 shows the next input symbol. Column 2 shows the new values of Low and High. Column 3 shows these values as scaled integers, after High has been decremented by 1. Column 4 shows the next digit sent to the output stream. Column 5 shows the new values of Low and High after being shifted to the left. Notice how the last step sends the four digits 3750 to the output stream. The final output is 717533750. Decoding is the opposite of encoding. We start with Low=0000, High=9999, and Code=7175 (the first four digits of the compressed stream). These are updated at each step of the decoding loop. Low and High approach each other (and both approach Code) until their most significant digits are the same. They are then shifted to the left, which separates them again, and Code is also shifted at that time. An index is calculated at each step and is used to search the cumulative frequencies column of Table 5.42 to figure out the current symbol. Each iteration of the loop consists of the following steps: 1. Calculate index:=((Code-Low+1)x10-1)/(High-Low+1) and truncate it to the nearest integer. (The number 10 is the total cumulative frequency in our example.) 2. Use index to find the next symbol by comparing it to the cumulative frequencies column in Table 5.42. In the example below, the first value of index is 7.1759, truncated to 7. Seven is between the 5 and the 10 in the table, so it selects the S. 3. Update Low and High according to Low:=Low+(High-Low+1)LowCumFreq[X]/10; High:=Low+(High-Low+1)HighCumFreq[X]/10-1; where LowCumFreq[X] and HighCumFreq[X] are the cumulative frequencies of symbol X and of the symbol above it in Table 5.42. 272 5. Statistical Methods 1 2 3 4 5 S L = 0+(1 - 0)Χ0.5= 0.5 5000 5000 H = 0+(1 - 0)Χ1.0= 1.0 9999 9999 W L = 0.5+(1 - .5)Χ0.4= 0.7 7000 7 0000 H = 0.5+(1 - .5)Χ0.5= 0.75 7499 7 4999 I L = 0+(0.5 - 0)Χ0.2= 0.1 1000 1 0000 H = 0+(0.5 - 0)Χ0.4= 0.2 1999 1 9999 S L = 0+(1 - 0)Χ0.5= 0.5 5000 5000 H = 0+(1 - 0)Χ1.0= 1.0 9999 9999 S L = 0.5+(1 - 0.5)Χ0.5= 0.75 7500 7500 H = 0.5+(1 - 0.5)Χ1.0= 1.0 9999 9999  L =0.75+(1 - 0.75)Χ0.0= 0.75 7500 7 5000 H =0.75+(1 - 0.75)Χ0.1= 0.775 7749 7 7499 M L = 0.5+(0.75 - 0.5)Χ0.1= 0.525 5250 5 2500 H = 0.5+(0.75 - 0.5)Χ0.2= 0.55 5499 5 4999 I L =0.25+(0.5 - 0.25)Χ0.2= 0.3 3000 3 0000 H =0.25+(0.5 - 0.25)Χ0.4= 0.35 3499 3 4999 S L = 0+(0.5 - 0)Χ0.5= .25 2500 2500 H = 0+(0.5 - 0)Χ1.0= 0.5 4999 4999 S L =0.25+(0.5 - 0.25)Χ0.5= 0.375 3750 3750 H =0.25+(0.5 - 0.25)Χ1.0= 0.5 4999 4999 Table 5.51: Encoding SWISSMISS by Shifting. 4. If the leftmost digits of Low and High are identical, shift Low, High, and Code one position to the left. Low gets a 0 entered on the right, High gets a 9, and Code gets the next input digit from the compressed stream. Here are all the decoding steps for our example: 0. Initialize Low=0000, High=9999, and Code=7175. 1. index= [(7175- 0 + 1) Χ 10 - 1]/(9999 - 0 + 1) = 7.1759 › 7. Symbol S is selected. Low = 0 +(9999-0+1)Χ5/10 = 5000. High = 0 +(9999-0+1)Χ10/10-1 = 9999. 2. index= [(7175 - 5000 + 1) Χ 10 - 1]/(9999 - 5000 + 1) = 4.3518 › 4. Symbol W is selected. Low = 5000+(9999-5000+1)Χ4/10 = 7000. High = 5000+(9999-5000+1)Χ5/10-1 = 7499. After the 7 is shifted out, we have Low=0000, High=4999, and Code=1753. 3. index= [(1753- 0 + 1) Χ 10 - 1]/(4999 - 0 + 1) = 3.5078 › 3. Symbol I is selected. Low = 0 + (4999 - 0 + 1) Χ 2/10 = 1000. High = 0 + (4999 - 0 + 1) Χ 4/10 - 1 = 1999. After the 1 is shifted out, we have Low=0000, High=9999, and Code=7533. 4. index= [(7533- 0 + 1) Χ 10 - 1]/(9999 - 0 + 1) = 7.5339 › 7. Symbol S is selected. Low = 0 +(9999-0+1)Χ5/10 = 5000. High = 0 +(9999-0+1)Χ10/10-1 = 9999. 5.9 Arithmetic Coding 273 5. index= [(7533 - 5000 + 1) Χ 10 - 1]/(9999 - 5000 + 1) = 5.0678 › 5. Symbol S is selected. Low = 5000+(9999-5000+1)Χ5/10 = 7500. High = 5000+(9999-5000+1)Χ10/10-1 = 9999. 6. index= [(7533 - 7500 + 1) Χ 10 - 1]/(9999 - 7500 + 1) = 0.1356 › 0. Symbol  is selected. Low = 7500+(9999-7500+1)Χ0/10 = 7500. High = 7500+(9999-7500+1)Χ1/10-1 = 7749. After the 7 is shifted out, we have Low=5000, High=7499, and Code=5337. 7. index= [(5337 - 5000 + 1) Χ 10 - 1]/(7499 - 5000 + 1) = 1.3516 › 1. Symbol M is selected. Low = 5000+(7499-5000+1)Χ1/10 = 5250. High = 5000+(7499-5000+1)Χ2/10-1 = 5499. After the 5 is shifted out we have Low=2500, High=4999, and Code=3375. 8. index= [(3375 - 2500 + 1) Χ 10 - 1]/(4999 - 2500 + 1) = 3.5036 › 3. Symbol I is selected. Low = 2500+(4999-2500+1)Χ2/10 = 3000. High = 2500+(4999-2500+1)Χ4/10-1 = 3499. After the 3 is shifted out we have Low=0000, High=4999, and Code=3750. 9. index= [(3750- 0 + 1) Χ 10 - 1]/(4999 - 0 + 1) = 7.5018 › 7. Symbol S is selected. Low = 0 +(4999-0+1)Χ5/10 = 2500. High = 0 +(4999-0+1)Χ10/10-1 = 4999. 10. index= [(3750 - 2500 + 1) Χ 10 - 1]/(4999 - 2500 + 1) = 5.0036 › 5. Symbol S is selected. Low = 2500+(4999-2500+1)Χ5/10 = 3750. High = 2500+(4999-2500+1)Χ10/10-1 = 4999.  Exercise 5.26: How does the decoder know to stop the loop at this point? 1 2 3 4 5 1 L=0+(1 - 0)Χ0.0 =0.0 000000 0 000000 H=0+(1 - 0)Χ0.023162= 0.023162 023162 0 231629 2 L=0+(0.231629 - 0)Χ0.0 =0.0 000000 0 000000 H=0+(0.231629 - 0)Χ0.023162= 0.00536478244 005364 0 053649 3 L=0+(0.053649 - 0)Χ0.0 =0.0 000000 0 000000 H=0+(0.053649 - 0)Χ0.023162= 0.00124261813 001242 0 012429 4 L=0+(0.012429 - 0)Χ0.0 =0.0 000000 0 000000 H=0+(0.012429 - 0)Χ0.023162= 0.00028788049 000287 0 002879 5 L=0+(0.002879 - 0)Χ0.0 =0.0 000000 0 000000 H=0+(0.002879 - 0)Χ0.023162= 0.00006668339 000066 0 000669 Table 5.52: Encoding a3a3a3a3a3 by Shifting. 274 5. Statistical Methods 5.9.2 Underflow Table 5.52 shows the steps in encoding the string a3a3a3a3a3 by shifting. This table is similar to Table 5.51, and it illustrates the problem of underflow. Low and High approach each other, and since Low is always 0 in this example, High loses its significant digits as it approaches Low. Underflow may happen not just in this case but in any case where Low and High need to converge very closely. Because of the finite size of the Low and High variables, they may reach values of, say, 499996 and 500003, and from there, instead of reaching values where their most significant digits are identical, they reach the values 499999 and 500000. Since the most significant digits are different, the algorithm will not output anything, there will not be any shifts, and the next iteration will only add digits beyond the first six ones. Those digits will be lost, and the first six digits will not change. The algorithm will iterate without generating any output until it reaches the eof. The solution to this problem is to detect such a case early and rescale both variables. In the example above, rescaling should be done when the two variables reach values of 49xxxx and 50yyyy. Rescaling should squeeze out the second most significant digits, end up with 4xxxx0 and 5yyyy9, and increment a counter cntr. The algorithm may have to rescale several times before the most-significant digits become equal. At that point, the most-significant digit (which can be either 4 or 5) should be output, followed by cntr zeros (if the two variables converged to 4) or nines (if they converged to 5). 5.9.3 Final Remarks All the examples so far have been in decimal, since the computations involved are easier to understand in this number base. It turns out that all the algorithms and rules described above apply to the binary case as well and can be used with only one change: Every occurrence of 9 (the largest decimal digit) should be replaced by 1 (the largest binary digit). The examples above don’t seem to show any compression at all. It seems that the three example strings SWISSMISS, a2a2a1a3a3, and a3a3a3a3eof are encoded into very long numbers. In fact, it seems that the length of the final code depends on the probabilities involved. The long probabilities of Table 5.46a generate long numbers in the encoding process, whereas the shorter probabilities of Table 5.42 result in the more reasonable Low and High values of Table 5.43. This behavior demands an explanation. I am ashamed to tell you to how many figures I carried these computations, having no other business at that time. —Isaac Newton To figure out the kind of compression achieved by arithmetic coding, we have to consider two facts: (1) In practice, all the operations are performed on binary numbers, so we have to translate the final results to binary before we can estimate the efficiency of the compression; (2) since the last symbol encoded is the eof, the final code does not have to be the final value of Low; it can be any value between Low and High. This makes it possible to select a shorter number as the final code that’s being output. Table 5.43 encodes the string SWISSMISS into the final Low and High values 0.71753375 and 0.717535. The approximate binary values of these numbers are 5.9 Arithmetic Coding 275 0.10110111101100000100101010111 and 0.1011011110110000010111111011, so we can select the number 10110111101100000100 as our final, compressed output. The ten-symbol string has thus been encoded into a 20-bit number. Does this represent good compression? The answer is yes. Using the probabilities of Table 5.42, it is easy to calculate the probability of the string SWISSMISS. It is P = 0.55Χ0.1Χ0.22Χ0.1Χ0.1 = 1.25Χ10-6. The entropy of this string is therefore -log2 P = 19.6096. Twenty bits are therefore the minimum needed in practice to encode the string. The symbols in Table 5.46a have probabilities 0.975, 0.001838, and 0.023162. These numbers require quite a few decimal digits, and as a result, the final Low and High values in Table 5.47 are the numbers 0.99462270125 and 0.994623638610941. Again it seems that there is no compression, but an analysis similar to the above shows compression that’s very close to the entropy. The probability of the string a2a2a1a3a3 is 0.9752Χ0.001838Χ0.0231622 . 9.37361Χ 10-7, and -log2 9.37361 Χ 10-7 . 20.0249. The binary representations of the final values of Low and High in Table 5.47 are 0.111111101001111110010111111001 and 0.111111101001111110100111101. We can select any number between these two, so we select 1111111010011111100, a 19-bit number. (This should have been a 21-bit number, but the numbers in Table 5.47 have limited precision and are not exact.)  Exercise 5.27: Given the three symbols a1, a2, and eof, with probabilities P1 = 0.4, P2 = 0.5, and Peof = 0.1, encode the string a2a2a2eof and show that the size of the final code equals the (practical) minimum. The following argument shows why arithmetic coding can, in principle, be a very efficient compression method. We denote by s a sequence of symbols to be encoded, and by b the number of bits required to encode it. As s gets longer, its probability P(s) gets smaller and b gets larger. Since the logarithm is the information function, it is easy to see that b should grow at the same rate that log2 P(s) shrinks. Their product should therefore be constant, or close to a constant. Information theory shows that b and P(s) satisfy the double inequality 2 . 2bP(s) < 4, which implies 1 - log2 P(s) . b < 2 - log2 P(s). (5.1) As s gets longer, its probability P(s) shrinks, the quantity -log2 P(s) becomes a large positive number, and the double inequality of Equation (5.1) shows that in the limit, b approaches -log2 P(s). This is why arithmetic coding can, in principle, compress a string of symbols to its theoretical limit. For more information on this topic, see [Moffat et al. 98] and [Witten et al. 87]. 276 5. Statistical Methods 5.10 Adaptive Arithmetic Coding Two features of arithmetic coding make it easy to extend: 1. One of the main encoding steps (page 266) updates NewLow and NewHigh. Similarly, one of the main decoding steps (step 3 on page 271) updates Low and High according to Low:=Low+(High-Low+1)LowCumFreq[X]/10; High:=Low+(High-Low+1)HighCumFreq[X]/10-1; This means that in order to encode symbol X, the encoder should be given the cumulative frequencies of the symbol and of the one above it (see Table 5.42 for an example of cumulative frequencies). This also implies that the frequency of X (or, equivalently, its probability) could be changed each time it is encoded, provided that the encoder and the decoder agree on how to do this. 2. The order of the symbols in Table 5.42 is unimportant. They can even be swapped in the table during the encoding process as long as the encoder and decoder do it in the same way. With this in mind, it is easy to understand how adaptive arithmetic coding works. The encoding algorithm has two parts: the probability model and the arithmetic encoder. The model reads the next symbol from the input stream and invokes the encoder, sending it the symbol and the two required cumulative frequencies. The model then increments the count of the symbol and updates the cumulative frequencies. The point is that the symbol’s probability is determined by the model from its old count, and the count is incremented only after the symbol has been encoded. This makes it possible for the decoder to mirror the encoder’s operations. The encoder knows what the symbol is even before it is encoded, but the decoder has to decode the symbol in order to find out what it is. The decoder can therefore use only the old counts when decoding a symbol. Once the symbol has been decoded, the decoder increments its count and updates the cumulative frequencies in exactly the same way as the encoder. The model should keep the symbols, their counts (frequencies of occurrence), and their cumulative frequencies in an array. This array should be kept in sorted order of the counts. Each time a symbol is read and its count is incremented, the model updates the cumulative frequencies, then checks to see whether it is necessary to swap the symbol with another one, to keep the counts in sorted order. It turns out that there is a simple data structure that allows for both easy search and update. This structure is a balanced binary tree housed in an array. (A balanced binary tree is a complete binary tree where some of the bottom-right nodes may be missing.) The tree should have a node for every symbol in the alphabet, and since it is balanced, its height is log2 n, where n is the size of the alphabet. For n = 256 the height of the balanced binary tree is 8, so starting at the root and searching for a node takes at most eight steps. The tree is arranged such that the most probable symbols (the ones with high counts) are located near the root, which speeds up searches. Table 5.53a shows an example of a ten-symbol alphabet with counts. Table 5.53b shows the same symbols sorted by count. The sorted array “houses” the balanced binary tree of Figure 5.55a. This is a simple, elegant way to build a tree. A balanced binary tree can be housed in an array without the use of any pointers. The rule is that the first array location (with index 1) houses 5.10 Adaptive Arithmetic Coding 277 a1 a2 a3 a4 a5 a6 a7 a8 a9 a10 11 12 12 2 5 1 2 19 12 8 (a) a8 a2 a3 a9 a1 a10 a5 a4 a7 a6 19 12 12 12 11 8 5 2 2 1 (b) Table 5.53: A Ten-Symbol Alphabet With Counts. the root, the two children of the node at array location i are housed at locations 2i and 2i + 1, and the parent of the node at array location j is housed at location j/2 . It is easy to see how sorting the array has placed the symbols with largest counts at and near the root. In addition to a symbol and its count, another value is now added to each tree node, the total counts of its left subtree. This will be used to compute cumulative frequencies. The corresponding array is shown in Table 5.54a. Assume that the next symbol read from the input stream is a9. Its count is incremented from 12 to 13. The model keeps the array in sorted order by searching for the farthest array element left of a9 that has a count smaller than that of a9. This search can be a straight linear search if the array is short enough, or a binary search if the array is long. In our case, symbols a9 and a2 should be swapped (Table 5.54b). Figure 5.55b shows the tree after the swap. Notice how the left-subtree counts have been updated. a8 a2 a3 a9 a1 a10 a5 a4 a7 a6 19 12 12 12 11 8 5 2 2 1 40 16 8 2 1 0 0 0 0 0 (a) a8 a9 a3 a2 a1 a10 a5 a4 a7 a6 19 13 12 12 11 8 5 2 2 1 41 16 8 2 1 0 0 0 0 0 (b) Tables 5.54: A Ten-Symbol Alphabet With Counts. Finally, here is how the cumulative frequencies are computed from this tree. When the cumulative frequency for a symbol X is needed, the model follows the tree branches from the root to the node containing X while adding numbers into an integer af. Each time a right branch is taken from an interior node N, af is incremented by the two numbers (the count and the left-subtree count) found in that node. When a left branch is taken, af is not modified. When the node containing X is reached, the left-subtree count of X is added to af, and af then contains the quantity LowCumFreq[X]. 278 5. Statistical Methods a4,2,0 a2,12,2 a9,13,16 a7,2,0 a6,1,0 a1,11,1 a10,8,0 a5,5,0 a3,12,8 a8,19,41 a4,2,0 a7,2,0 a6,1,0 a5 a10,8,0 ,5,0 a9,12,2 a1,11,1 a2,12,16 a8,19,40 a3,12,8 (a) (b) a4 2 0—1 a9 12 2—13 a7 2 14—15 a2 12 16—27 a6 1 28—28 a1 11 29—39 a8 19 40—58 a10 8 59—66 a3 12 67—78 a5 5 79—83 (c) Figure 5.55: Adaptive Arithmetic Coding. 5.10 Adaptive Arithmetic Coding 279 As an example, we trace the tree of Figure 5.55a from the root to symbol a6, whose cumulative frequency is 28. A right branch is taken at node a2, adding 12 and 16 to af. A left branch is taken at node a1, adding nothing to af. When reaching a6, its left-subtree count, 0, is added to af. The result in af is 12+16 = 28, as can be verified from Figure 5.55c. The quantity HighCumFreq[X] is obtained by adding the count of a6 (which is 1) to LowCumFreq[X]. To trace the tree and find the path from the root to a6, the algorithm performs the following steps: 1. Find a6 in the array housing the tree by means of a binary search. In our example the node with a6 is found at array location 10. 2. Integer-divide 10 by 2. The remainder is 0, which means that a6 is the left child of its parent. The quotient is 5, which is the array location of the parent. 3. Location 5 of the array contains a1. Integer-divide 5 by 2. The remainder is 1, which means that a1 is the right child of its parent. The quotient is 2, which is the array location of a1’s parent. 4. Location 2 of the array contains a2. Integer-divide 2 by 2. The remainder is 0, which means that a2 is the left child of its parent. The quotient is 1, the array location of the root, so the process stops. The PPM compression method, Section 5.14, is a good example of a statistical model that invokes an arithmetic encoder in the way described here. The driver held out a letter. Boldwood seized it and opened it, expecting another anonymous one—so greatly are people’s ideas of probability a mere sense that precedent will repeat itself. “I don’t think it is for you, sir,” said the man, when he saw Boldwood’s action. “Though there is no name I think it is for your shepherd.” —Thomas Hardy, Far From The Madding Crowd 5.10.1 Range Encoding The use of integers in arithmetic coding is a must in any practical implementation, but it results in slow encoding because of the need for frequent renormalizations. The main steps in any integer-based arithmetic coding implementation are (1) proportional range reduction and (2) range expansion (renormalization). Range encoding (or range coding) is an improvement to arithmetic coding that reduces the number of renormalizations and thereby speeds up integer-based arithmetic coding by factors of up to 2. The main references are [Schindler 98] and [Campos 06], and the description here is based on the former. The main idea is to treat the output not as a binary number, but as a number to another base (256 is commonly used as a base, implying that each digit is a byte). This requires fewer renormalizations and no bitwise operations. The following analysis may shed light on this method. At any point during arithmetic coding, the output consists of four parts as follows: 1. The part already written on the output. This part will not change. 2. One digit (bit, byte, or a digit to another base) that may be modified by at most one carry when adding to the lower end of the interval. (There cannot be two carries 280 5. Statistical Methods because when this digit was originally determined, the range was less than or equal to one unit. Two carries require a range greater than one unit.) 3. A (possibly empty) block of digits that passes on a carry (1 in binary, 9 in decimal, 255 for base-256, etc.) and are represented by a counter counting their number. 4. The low variable of the encoder. The following states can occur while data is encoded: No renormalization is needed because the range is in the desired interval. The low end plus the range (this is the upper end of the interval) will not produce any carry. In this case the second and third parts can be output because they will never change. The digit produced will become part two, and part three will be empty. The low end has already produced a carry. In this case, the (modified) second and third parts can be output; there will not be another carry. Set the second and third part as before. The digit produced will pass on a possible future carry, so it is added to the block of digits of part three. The difference between conventional integer-based arithmetic coding and range coding is that in the latter, part two, which may be modified by a carry, has to be stored explicitly. With binary output this part is always 0 since the 1’s are always added to the carry-passing-block. Implementing that is straightforward. More information and code can be found in [Campos 06]. Range coding is used in LZMA (Section 6.26). 5.11 The QM Coder JPEG (Section 7.10) is an important image compression method. It uses arithmetic coding, but not in the way described in Section 5.9. The arithmetic coder of JPEG is called the QM-coder and is described in this section. It is designed for simplicity and speed, so it is limited to input symbols that are single bits and it uses an approximation instead of a multiplication. It also uses fixed-precision integer arithmetic, so it has to resort to renormalization of the probability interval from time to time, in order for the approximation to remain close to the true multiplication. For more information on this method, see [IBM 88], [Pennebaker and Mitchell 88a], and [Pennebaker and Mitchell 88b]. A slight confusion arises because the arithmetic coder of JPEG 2000 (Section 8.19) and JBIG2 (Section 7.15) is called the MQ-coder and is not the same as the QM-coder (we are indebted to Christopher M. Brislawn for pointing this out).  Exercise 5.28: The QM-coder is limited to input symbols that are single bits. Suggest a way to convert an arbitrary set of symbols to a stream of bits. The main idea behind the QM-coder is to classify each input symbol (which is a single bit) as either the more probable symbol (MPS) or the less probable symbol (LPS). Before the next bit is input, the QM-encoder uses a statistical model to determine 5.11 The QM Coder 281 whether a 0 or a 1 is more probable at that point. It then inputs the next bit and classifies it according to its actual value. If the model predicts, for example, that a 0 is more probable, and the next bit turns out to be a 1, the encoder classifies it as an LPS. It is important to understand that the only information encoded in the compressed stream is whether the next bit is MPS or LPS. When the stream is decoded, all that the decoder knows is whether the bit that has just been decoded is an MPS or an LPS. The decoder has to use the same statistical model to determine the current relation between MPS/LPS and 0/1. This relation changes, of course, from bit to bit, since the model is updated identically (in lockstep) by the encoder and decoder each time a bit is input by the former or decoded by the latter. The statistical model also computes a probability Qe for the LPS, so the probability of the MPS is (1 - Qe). Since Qe is the probability of the less probable symbol, it is in the range [0, 0.5]. The encoder divides the probability interval A into two subintervals according to Qe and places the LPS subinterval (whose size is AΧQe) above the MPS subinterval [whose size is A(1 - Qe)], as shown in Figure 5.56b. Notice that the two subintervals in the figure are closed at the bottom and open at the top. This should be compared with the way a conventional arithmetic encoder divides the same interval (Figure 5.56a, where the numbers are taken from Table 5.43). (a) 0 0 LPS subinterval MPS subinterval 1 1 0.5 0.5 0.7 0.7 0.71 0.71 0.72 0.715 0.72 0.75 0.75 (b) A(1-Qe) A(1-Qe) AΧQe A Figure 5.56: Division of the Probability Interval. In conventional arithmetic coding, the interval is narrowed all the time, and the final output is any number inside the final subinterval. In the QM-coder, for simplicity, each step adds the bottom of the selected subinterval to the output-so-far. We denote the output string by C. If the current bit read from the input is the MPS, the bottom of the MPS subinterval (i.e., the number 0) is added to C. If the current bit is the LPS, the bottom of the LPS subinterval [i.e., the number A(1 - Qe)] is added to C. After C is updated in this way, the current probability interval A is shrunk to the size of the selected subinterval. The probability interval is always in the range [0,A), and A gets smaller at each step. This is the main principle of the QM-encoder, and it is expressed by the rules After MPS: C is unchanged, A ‹ A(1 - Qe), After LPS: C ‹ C + A(1 - Qe), A‹ AΧQe. (5.2) 282 5. Statistical Methods These rules set C to point to the bottom of the MPS or the LPS subinterval, depending on the classification of the current input bit. They also set A to the new size of the subinterval. Table 5.57 lists the values of A and C when four symbols, each a single bit, are encoded. We assume that they alternate between an LPS and an MPS and that Qe = 0.5 for all four steps (normally, of course, the statistical model yields different values of Qe all the time). It is easy to see how the probability interval A shrinks from 1 to 0.0625, and how the output C grows from 0 to 0.625. Table 5.59 is similar, but uses Qe = 0.1 for all four steps. Again A shrinks, to 0.0081, and C grows, to 0.981. Figures 5.58 and 5.60 illustrate graphically the division of the probability interval A into an LPS and an MPS.  Exercise 5.29: Repeat these calculations for the case where all four symbols are LPS and Qe = 0.5, then for the case where they are MPS and Qe = 0.1. The principle of the QM-encoder is simple and easy to understand, but it involves two problems. The first is the fact that the interval A, which starts at 1, shrinks all the time and requires high precision to distinguish it from zero. The solution to this problem is to maintain A as an integer and double it every time it gets too small. This is called renormalization. It is fast, since it is done by a logical left shift; no multiplication is needed. Each time A is doubled, C is also doubled. The second problem is the multiplication AΧQe used in subdividing the probability interval A. A fast compression method should avoid multiplications and divisions and should try to replace them with additions, subtractions, and shifts. It turns out that the second problem is also solved by renormalization. The idea is to keep the value of A close to 1, so that Qe will not be very different from the product A Χ Qe. The multiplication is approximated by Qe. How can we use renormalization to keep A close to 1? The first idea that comes to mind is to double A when it gets just a little below 1, say to 0.9. The problem is that doubling 0.9 yields 1.8, closer to 2 than to 1. If we let A get below 0.5 before doubling it, the result will be less than 1. It does not take long to realize that 0.75 is a good minimum value for renormalization. If A reaches this value at a certain step, it is doubled, to 1.5. If it reaches a smaller value, such as 0.6 or 0.55, it ends up even closer to 1 when doubled. If A reaches a value less than 0.5 at a certain step, it has to be renormalized by doubling it several times, each time also doubling C. An example is the second row of Table 5.59, where A shrinks from 1 to 0.1 in one step, because of a very small probability Qe. In this case, A has to be doubled three times, from 0.1 to 0.2, to 0.4, to 0.8, in order to bring it into the desired range [0.75, 1.5). We conclude that A can go down to 0 (or very close to 0) and can be at most 1.5 (actually, less than 1.5, since our intervals are always open at the high end).  Exercise 5.30: In what case does A always have to be renormalized? Approximating the multiplication A Χ Qe by Qe changes the main rules of the QM-encoder to After MPS: C is unchanged, A ‹ A(1 - Qe) . A - Qe, After LPS: C ‹ C + A(1 - Qe) . C + A - Qe, A ‹ AΧQe . Qe. 5.11 The QM Coder 283 Symbol C A Initially 0 1 s1 (LPS) 0 + 1(1 - 0.5) = 0.5 1Χ0.5 = 0.5 s2 (MPS) unchanged 0.5Χ(1 - 0.5) = 0.25 s3 (LPS) 0.5 + 0.25(1 - 0.5) = 0.625 0.25Χ0.5 = 0.125 s4 (MPS) unchanged 0.125Χ(1 - 0.5) = 0.0625 Table 5.57: Encoding Four Symbols With Qe = 0.5. 0 0 1 0.5 0.5 A(1-Qe) 0.25 0 0.25 0.125 0 0.125 0.0625 0.0625 0 Figure 5.58: Division of the Probability Interval. Symbol C A Initially 0 1 s1 (LPS) 0 + 1(1 - 0.1) = 0.9 1Χ0.1 = 0.1 s2 (MPS) unchanged 0.1Χ(1 - 0.1) = 0.09 s3 (LPS) 0.9 + 0.09(1 - 0.1) = 0.981 0.09Χ0.1 = 0.009 s4 (MPS) unchanged 0.009Χ(1 - 0.1) = 0.0081 Table 5.59: Encoding Four Symbols With Qe = 0.1. 0 0 1 0.9 0.1 0.09 A(1-Qe) A 0 0.09 0.081 0 0.009 0.0081 0 0.0081 Figure 5.60: Division of the Probability Interval. 284 5. Statistical Methods In order to include renormalization in these rules, we have to choose an integer representation for A where real values in the range [0, 1.5) are represented as integers. Since many current and old computers have 16-bit words, it makes sense to choose a representation where 0 is represented by a word of 16 zero bits and 1.5 is represented by the smallest 17-bit number, which is 216 = 6553610 = 1000016 = 10. . . 0    16 2. This way we can represent 65536 real values in the range [0, 1.5) as 16-bit integers, where the largest 16-bit integer, 65535, represents a real value slightly less than 1.5. Here are a few important examples of such values: 0.75 = 1.5/2 = 215 = 3276810 = 800016, 1 = 0.75(4/3) = 4369010 = AAAA16, 0.5 = 43690/2 = 2184510 = 555516, 0.25 = 21845/2 = 1092310 = 2AAB16. (The optimal value of 1 in this representation is AAAA16, but the way we associate the real values of A with the 16-bit integers is somewhat arbitrary. The important thing about this representation is to achieve accurate interval subdivision, and the subdivision is done by either A ‹ A - Qe or A ‹ Qe. The accuracy of the subdivision depends, therefore, on the relative values of A and Qe, and it has been found experimentally that the average value of A is B55A16, so this value, instead of AAAA16, is associated in the JPEG QM-coder with A = 1. The difference between the two values AAAA and B55A is AB016 = 273610. The JBIG QM-coder uses a slightly different value for 1.) Renormalization can now be included in the main rules of the QM-encoder, which become After MPS: C is unchanged, A ‹ A - Qe, if A < 800016 renormalize A and C. After LPS: C ‹ C + A - Qe, A ‹ Qe, renormalize A and C. (5.3) Tables 5.61 and 5.62 list the results of applying these rules to the examples shown in Tables 5.57 and 5.59, respectively.  Exercise 5.31: Repeat these calculations with renormalization for the case where all four symbols are LPS and Qe = 0.5. Following this, repeat the calculations for the case where they are all MPS and Qe = 0.1. (Compare this exercise with Exercise 5.29.) The next point that has to be considered in the design of the QM-encoder is the problem of interval inversion. This is the case where the size of the subinterval allocated to the MPS becomes smaller than the LPS subinterval. This problem may occur when Qe is close to 0.5 and is a result of the approximation to the multiplication. It is illustrated in Table 5.63, where four MPS symbols are encoded with Qe = 0.45. In the third row of the table the interval A is doubled from 0.65 to 1.3. In the fourth row it is reduced to 0.85. This value is greater than 0.75, so no renormalization takes place; yet the subinterval allocated to the MPS becomes A - Qe = 0.85 - 0.45 = 0.40, which is smaller than the LPS subinterval, which is Qe = 0.45. Clearly, the problem occurs when Qe > A/2, a relation that can also be expressed as Qe > A - Qe. 5.11 The QM Coder 285 Symbol C A Renor. A Renor. C Initially 0 1 s1 (LPS) 0 + 1 - 0.5 = 0.5 0.5 1 1 s2 (MPS) unchanged 1 - 0.5 = 0.5 1 2 s3 (LPS) 2 + 1 - 0.5 = 2.5 0.5 1 5 s4 (MPS) unchanged 1 - 0.5 = 0.5 1 10 Table 5.61: Renormalization Added to Table 5.57. Symbol C A Renor. A Renor. C Initially 0 1 s1 (LPS) 0 + 1 - 0.1 = 0.9 0.1 0.8 0.9·23 = 7.2 s2 (MPS) unchanged 7.2 0.8 - 0.1 = 0.7 1.4 7.2·2 = 14.4 s3 (LPS) 14.4 + 1.4 - 0.1 = 15.7 0.1 0.8 15.7·23 = 125.6 s4 (MPS) unchanged 0.8 - 0.1 = 0.7 1.4 125.6·2 = 251.2 Table 5.62: Renormalization Added to Table 5.59. SymbolC ARenor. A Renor. C Initially 0 1 s1 (MPS) 0 1 - 0.45 = 0.55 1.1 0 s2 (MPS) 0 1.1 - 0.45 = 0.65 1.3 0 s3 (MPS) 0 1.3 - 0.45 = 0.85 s4 (MPS) 0 0.85 - 0.45 = 0.40 0.8 0 Table 5.63: Illustrating Interval Inversion. The solution is to interchange the two subintervals whenever the LPS subinterval becomes greater than the MPS subinterval. This is called conditional exchange. The condition for interval inversion is Qe > A - Qe, but since Qe . 0.5, we get A - Qe < Qe . 0.5, and it becomes obvious that both Qe and A - Qe (i.e., both the LPS and MPS subintervals) are less than 0.75, so renormalization must take place. This is why the test for conditional exchange is performed only after the encoder has decided that renormalization is needed. The new rules for the QM-encoder are shown in Figure 5.64. The QM-Decoder: The QM-decoder is the reverse of the QM-encoder. For simplicity we ignore renormalization and conditional exchange, and we assume that the QM-encoder operates by the rules of Equation (5.2). Reversing the way C is updated in those rules yields the rules for the QM-decoder (the interval A is updated in the same way): After MPS: C is unchanged, A ‹ A(1 - Qe), After LPS: C ‹ C - A(1 - Qe), A‹ AΧQe. (5.4) These rules are demonstrated using the data of Table 5.57. The four decoding steps are as follows: 286 5. Statistical Methods After MPS: C is unchanged A ‹A - Qe; % The MPS subinterval if A < 800016 then % if renormalization needed if A < Qe then % if inversion needed C ‹C + A; % point to bottom of LPS A ‹Qe % Set A to LPS subinterval endif; renormalize A and C; endif; After LPS: A ‹A - Qe; % The MPS subinterval if A . Qe then % if interval sizes not inverted C ‹C + A; % point to bottom of LPS A ‹Qe % Set A to LPS subinterval endif; renormalize A and C; Figure 5.64: QM-Encoder Rules With Interval Inversion. Step 1: C = 0.625, A = 1, the dividing line is A(1-Qe) = 1(1-0.5) = 0.5, so the LPS and MPS subintervals are [0, 0.5) and [0.5, 1). Since C points to the upper subinterval, an LPS is decoded. The new C is 0.625-1(1-0.5) = 0.125 and the new A is 1Χ0.5 = 0.5. Step 2: C = 0.125, A = 0.5, the dividing line is A(1 - Qe) = 0.5(1 - 0.5) = 0.25, so the LPS and MPS subintervals are [0, 0.25) and [0.25, 0.5), and an MPS is decoded. C is unchanged, and the new A is 0.5(1 - 0.5) = 0.25. Step 3: C = 0.125, A = 0.25, the dividing line is A(1 - Qe) = 0.25(1 - 0.5) = 0.125, so the LPS and MPS subintervals are [0, 0.125) and [0.125, 0.25), and an LPS is decoded. The new C is 0.125 - 0.25(1 - 0.5) = 0, and the new A is 0.25 Χ 0.5 = 0.125. Step 4: C = 0, A = 0.125, the dividing line is A(1 - Qe) = 0.125(1 - 0.5) = 0.0625, so the LPS and MPS subintervals are [0, 0.0625) and [0.0625, 0.125), and an MPS is decoded. C is unchanged, and the new A is 0.125(1 - 0.5) = 0.0625.  Exercise 5.32: Use the rules of Equation (5.4) to decode the four symbols encoded in Table 5.59. Probability Estimation: The QM-encoder uses a novel, interesting, and littleunderstood method for estimating the probability Qe of the LPS. The first method that comes to mind in trying to estimate the probability of the next input bit is to initialize Qe to 0.5 and update it by counting the numbers of zeros and ones that have been input so far. If, for example, 1000 bits have been input so far, and 700 of them were zeros, then 0 is the current MPS, with probability 0.7, and the probability of the LPS is Qe = 0.3. Notice that Qe should be updated before the next input bit is read and encoded, since otherwise the decoder would not be able to mirror this operation (the decoder does not know what the next bit is). This method produces good results, but is 5.11 The QM Coder 287 slow, since Qe should be updated often (ideally, for each input bit), and the calculation involves a division (dividing 700/1000 in our example). The method used by the QM-encoder is based on a table of preset Qe values. Qe is initialized to 0.5 and is modified when renormalization takes place, not for every input bit. Table 5.65 illustrates the process. The Qe index is initialized to zero, so the first value of Qe is 0AC116 or very close to 0.5. After the first MPS renormalization, the Qe index is incremented by 1, as indicated by column “Incr MPS.” A Qe index of 1 implies a Qe value of 0A8116 or 0.49237, slightly smaller than the original, reflecting the fact that the renormalization occurred because of an MPS. If, for example, the current Qe index is 26, and the next renormalization is LPS, the index is decremented by 3, as indicated by column “Decr LPS,” reducing Qe to 0.00421. The method is not applied very often, and it involves only table lookups and incrementing or decrementing the Qe index: fast, simple operations. Qe Hex Dec Decr Incr MPS Qe Hex Dec Decr Incr MPS index Qe Qe LPS MPS exch index Qe Qe LPS MPS exch 0 0AC1 0.50409 0 1 1 15 0181 0.07050 2 1 0 1 0A81 0.49237 1 1 0 16 0121 0.05295 2 1 0 2 0A01 0.46893 1 1 0 17 00E1 0.04120 2 1 0 3 0901 0.42206 1 1 0 18 00A1 0.02948 2 1 0 4 0701 0.32831 1 1 0 19 0071 0.02069 2 1 0 5 0681 0.30487 1 1 0 20 0059 0.01630 2 1 0 6 0601 0.28143 1 1 0 21 0053 0.01520 2 1 0 7 0501 0.23456 2 1 0 22 0027 0.00714 2 1 0 8 0481 0.21112 2 1 0 23 0017 0.00421 2 1 0 9 0441 0.19940 2 1 0 24 0013 0.00348 3 1 0 10 0381 0.16425 2 1 0 25 000B 0.00201 2 1 0 11 0301 0.14081 2 1 0 26 0007 0.00128 3 1 0 12 02C1 0.12909 2 1 0 27 0005 0.00092 2 1 0 13 0281 0.11737 2 1 0 28 0003 0.00055 3 1 0 14 0241 0.10565 2 1 0 29 0001 0.00018 2 0 0 Table 5.65: Probability Estimation Table (Illustrative). The column labeled “MPS exch” in Table 5.65 contains the information for the conditional exchange of the MPS and LPS definitions at Qe = 0.5. The zero value at the bottom of column “Incr MPS” should also be noted. If the Qe index is 29 and an MPS renormalization occurs, this zero causes the index to stay at 29 (corresponding to the smallest Qe value). Table 5.65 is used here for illustrative purposes only. The JPEG QM-encoder uses Table 5.66, which has the same format but is harder to understand, since its Qe values are not listed in sorted order. This table was prepared using probability estimation concepts based on Bayesian statistics. We now justify this probability estimation method with an approximate calculation that suggests that the Qe values obtained by this method will adapt to and closely approach the correct LPS probability of the binary input stream. The method updates 288 5. Statistical Methods Qe each time a renormalization occurs, and we know, from Equation (5.3), that this happens every time an LPS is input, but not for all MPS values. We therefore imagine an ideal balanced input stream where for each LPS bit there is a sequence of consecutive MPS bits. We denote the true (but unknown) LPS probability by q, and we try to show that the Qe values produced by the method for this ideal case are close to q. Equation (5.3) lists the main rules of the QM-encoder and shows how the probability interval A is decremented by Qe each time an MPS is input and encoded. Imagine a renormalization that brings A to a value A1 (between 1 and 1.5), followed by a sequence of N consecutive MPS bits that reduce A in steps of Qe from A1 to a value A2 that requires another renormalization (i.e., A2 is less than 0.75). It is clear that N = .A Qe !, where .A = A1 - A2. Since q is the true probability of an LPS, the probability of having N MPS bits in a row is P = (1 - q)N. This implies ln P = N ln(1 - q), which, for a small q, can be approximated by ln P . N(-q) = - .A Qe q, or P . exp- .A Qe q . (5.5) Since we are dealing with an ideal balanced input stream, we are interested in the value P = 0.5, because it implies equal numbers of LPS and MPS renormalizations. From P = 0.5 we get ln P = -ln 2, which, when combined with Equation (5.5), yields Qe = .A ln 2 q. This is fortuitous because ln 2 . 0.693 and .A is typically a little less than 0.75. We can say that for our ideal balanced input stream, Qe . q, providing one justification for our estimation method. Another justification is provided by the way P depends on Qe [shown in Equation (5.5)]. If Qe gets larger than q, P also gets large, and the table tends to move to smaller Qe values. In the opposite case, the table tends to select larger Qe values. 5.11 The QM Coder 289 Qe Hex Next-Index MPS Qe Hex Next-Index MPS index Qe LPS MPS exch index Qe LPS MPS exch 0 5A1D 1 1 1 57 01A4 55 58 0 1 2586 14 2 0 58 0160 56 59 0 2 1114 16 3 0 59 0125 57 60 0 3 080B 18 4 0 60 00F6 58 61 0 4 03D8 20 5 0 61 00CB 59 62 0 5 01DA 23 6 0 62 00AB 61 63 0 6 00E5 25 7 0 63 008F 61 32 0 7 006F 28 8 0 64 5B12 65 65 1 8 0036 30 9 0 65 4D04 80 66 0 9 001A 33 10 0 66 412C 81 67 0 10 000D 35 11 0 67 37D8 82 68 0 11 0006 9 12 0 68 2FE8 83 69 0 12 0003 10 13 0 69 293C 84 70 0 13 0001 12 13 0 70 2379 86 71 0 14 5A7F 15 15 1 71 1EDF 87 72 0 15 3F25 36 16 0 72 1AA9 87 73 0 16 2CF2 38 17 0 73 174E 72 74 0 17 207C 39 18 0 74 1424 72 75 0 18 17B9 40 19 0 75 119C 74 76 0 19 1182 42 20 0 76 0F6B 74 77 0 20 0CEF 43 21 0 77 0D51 75 78 0 21 09A1 45 22 0 78 0BB6 77 79 0 22 072F 46 23 0 79 0A40 77 48 0 23 055C 48 24 0 80 5832 80 81 1 24 0406 49 25 0 81 4D1C 88 82 0 25 0303 51 26 0 82 438E 89 83 0 26 0240 52 27 0 83 3BDD 90 84 0 27 01B1 54 28 0 84 34EE 91 85 0 28 0144 56 29 0 85 2EAE 92 86 0 29 00F5 57 30 0 86 299A 93 87 0 30 00B7 59 31 0 87 2516 86 71 0 31 008A 60 32 0 88 5570 88 89 1 32 0068 62 33 0 89 4CA9 95 90 0 33 004E 63 34 0 90 44D9 96 91 0 34 003B 32 35 0 91 3E22 97 92 0 35 002C 33 9 0 92 3824 99 93 0 36 5AE1 37 37 1 93 32B4 99 94 0 37 484C 64 38 0 94 2E17 93 86 0 38 3A0D 65 39 0 95 56A8 95 96 1 39 2EF1 67 40 0 96 4F46 101 97 0 40 261F 68 41 0 97 47E5 102 98 0 41 1F33 69 42 0 98 41CF 103 99 0 42 19A8 70 43 0 99 3C3D 104 100 0 43 1518 72 44 0 100 375E 99 93 0 44 1177 73 45 0 101 5231 105 102 0 45 0E74 74 46 0 102 4C0F 106 103 0 46 0BFB 75 47 0 103 4639 107 104 0 47 09F8 77 48 0 104 415E 103 99 0 48 0861 78 49 0 105 5627 105 106 1 49 0706 79 50 0 106 50E7 108 107 0 50 05CD 48 51 0 107 4B85 109 103 0 51 04DE 50 52 0 108 5597 110 109 0 52 040F 50 53 0 109 504F 111 107 0 53 0363 51 54 0 110 5A10 110 111 1 54 02D4 52 55 0 111 5522 112 109 0 55 025C 53 56 0 112 59EB 112 111 1 56 01F8 54 57 0 Table 5.66: The QM-Encoder Probability Estimation Table. 290 5. Statistical Methods 5.12 Text Compression Before delving into the details of the next method, here is a general discussion of text compression. Most text compression methods are either statistical or dictionary based. The latter class breaks the text into fragments that are saved in a data structure called a dictionary. When a fragment of new text is found to be identical to one of the dictionary entries, a pointer to that entry is written on the compressed stream, to become the compression of the new fragment. The former class, on the other hand, consists of methods that develop statistical models of the text. A common statistical method consists of a modeling stage followed by a coding stage. The model assigns probabilities to the input symbols, and the coding stage actually codes the symbols based on those probabilities. The model can be static or dynamic (adaptive). Most models are based on one of the following two approaches. Frequency: The model assigns probabilities to the text symbols based on their frequencies of occurrence, such that commonly-occurring symbols are assigned short codes. A static model uses fixed probabilities, whereas a dynamic model modifies the probabilities “on the fly” while text is being input and compressed. Context: The model considers the context of a symbol when assigning it a probability. Since the decoder does not have access to future text, both encoder and decoder must limit the context to past text, i.e., to symbols that have already been input and processed. In practice, the context of a symbol is the N symbols preceding it (where N is a parameter). We therefore say that a context-based text compression method uses the context of a symbol to predict it (i.e., to assign it a probability). Technically, such a method is said to use an “order-N” Markov model. The PPM method, Section 5.14, is an excellent example of a context-based compression method, although the concept of context can also be used to compress images. Some modern context-based text compression methods perform a transformation on the input data and then apply a statistical model to assign probabilities to the transformed symbols. Good examples of such methods are the Burrows-Wheeler method, Section 11.1, also known as the Burrows-Wheeler transform, or block sorting; the technique of symbol ranking, Section 11.2; and the ACB method, Section 11.3, which uses an associative dictionary. Reference [Bell et al. 90] is an excellent introduction to text compression. It also describes many important algorithms and approaches to this important problem. 5.13 The Hutter Prize Among the different types of digital data, text is the smallest in the sense that text files are much smaller than audio, image, or video files. A typical 300–400-page book may contain about 250,000 words or about a million characters, so it fits in a 1 MB file. A raw (uncompressed) 1,024 Χ 1,024 color image, on the other hand, contains one mega pixels and therefore becomes a 3 MB file (three bytes for the color components of each pixel). Video files are much bigger. This is why many workers in the compression field concentrate their efforts on video and image compression, but there is a group of “hard core” researchers who are determined to squeeze out the last bit of redundancy from 5.13 The Hutter Prize 291 text and compress text all the way down to its entropy. In order to encourage this group of enthusiasts, scientists, and programmers, Marcus Hutter decided to offer the prize that now bears his name. In August 2006, Hutter selected a specific 100-million-character text file with text from the English wikipedia, named it enwik8, and announced a prize with initial funding of 50,000 euros [Hutter 08]. Two other organizers of the prize are Matt Mahoney and Jim Bowery. Specifically, the prize awards 500 euros for each 1% improvement in the compression of enwik8. The following is a quotation from [wikiHutter 08]. The goal of the Hutter Prize is to encourage research in artificial intelligence (AI). The organizers believe that text compression and AI are equivalent problems. Hutter proved that the optimal behavior of a goal seeking agent in an unknown but computable environment is to guess at each step that the environment is controlled by the shortest program consistent with all interaction so far. Unfortunately, there is no general solution because Kolmogorov complexity is not computable. Hutter proved that in the restricted case (called AIXItl) where the environment is restricted to time t and space l, a solution can be computed in time O(t Χ 2l), which is still intractable. The organizers further believe that compressing natural language text is a hard AI problem, equivalent to passing the Turing test. Thus, progress toward one goal represents progress toward the other. They argue that predicting which characters are most likely to occur next in a text sequence requires vast realworld knowledge. A text compressor must solve the same problem in order to assign the shortest codes to the most likely text sequences. Anyone can participate in this competition. A competitor has to submit either an encoder and decoder or a compressed enwik8 and a decoder (there are certain time and memory constraints on the decoder). The source code is not required and the compression algorithm does not have to be general; it may be optimized for enwik8. The total size of the compressed enwik8 plus the decoder should not exceed 99% of the previous winning entry. For each step of 1% improvement, the lucky winner receives 500 euros. Initially, the compression baseline was the impressive 18,324,887 bytes, representing a compression ratio of 0.183 (achieved by PAQ8F, Section 5.15). So far, only Matt Mahoney and Alexander Ratushnyak, won prizes for improvements to the baseline, with Ratushnyak winning twice! The awards are computed by the simple expression Z(L - S)/L, where Z is the total prize funding (starting at 50,000 euros and possibly increasing in the future), S is the new record (size of the encoder or size of the compressed file plus the decoder), and L is the previous record. The minimum award is 3% of Z. 292 5. Statistical Methods 5.14 PPM The PPM method is a sophisticated, state of the art compression method originally developed by J. Cleary and I. Witten [Cleary and Witten 84], with extensions and an implementation by A. Moffat [Moffat 90]. The method is based on an encoder that maintains a statistical model of the text. The encoder inputs the next symbol S, assigns it a probability P, and sends S to an adaptive arithmetic encoder, to be encoded with probability P. The simplest statistical model counts the number of times each symbol has occurred in the past and assigns the symbol a probability based on that. Assume that 1217 symbols have been input and encoded so far, and 34 of them were the letter q. If the next symbol is a q, it is assigned a probability of 34/1217 and its count is incremented by 1. The next time q is seen, it will be assigned a probability of 35/t, where t is the total number of symbols input up to that point (not including the last q). The next model up is a context-based statistical model. The idea is to assign a probability to symbol S depending not just on the frequency of the symbol but on the contexts in which it has occurred so far. The letter h, for example, occurs in “typical” English text (Table Intro.1) with a probability of about 5%. On average, we expect to see an h about 5% of the time. However, if the current symbol is t, there is a high probability (about 30%) that the next symbol will be h, since the digram th is common in English. We say that the model of typical English predicts an h in such a case. If the next symbol is in fact h, it is assigned a large probability. In cases where an h is the second letter of an unlikely digram, say xh, the h is assigned a smaller probability. Notice that the word “predicts” is used here to mean “estimate the probability of.” A similar example is the letter u, which has a probability of about 2%. When a q is encountered, however, there is a probability of more than 99% that the next letter will be a u.  Exercise 5.33: We know that in English, a q must be followed by a u. Why not just say that the probability of the digram qu is 100%? A static context-based modeler always uses the same probabilities. It contains static tables with the probabilities of all the possible digrams (or trigrams) of the alphabet and uses the tables to assign a probability to the next symbol S depending on the symbol (or, in general, on the context) C preceding it. We can imagine S and C being used as indices for a row and a column of a static frequency table. The table itself can be constructed by accumulating digram or trigram frequencies from large quantities of text. Such a modeler is simple and produces good results on average, but has two problems. The first is that some input streams may be statistically very different from the data originally used to prepare the table. A static encoder may create considerable expansion in such a case. The second problem is zero probabilities. What if after reading and analyzing huge amounts of English text, we still have never encountered the trigram qqz? The cell corresponding to qqz in the trigram frequency table will contain zero. The arithmetic encoder, Sections 5.9 and 5.10, requires all symbols to have nonzero probabilities. Even if a different encoder, such as Huffman, is used, all the probabilities involved must be nonzero. (Recall that the Huffman method works by combining two low-probability symbols into one high-probability symbol. If 5.14 PPM 293 two zero-probability symbols are combined, the resulting symbol will have the same zero probability.) Another reason why a symbol must have nonzero probability is that its entropy (the smallest number of bits into which it can be encoded) depends on log2 P, which is undefined for P = 0 (but gets very large when P › 0). This zero-probability problem faces any model, static or adaptive, that uses probabilities of occurrence of symbols to achieve compression. Two simple solutions are traditionally adopted for this problem, but neither has any theoretical justification. 1. After analyzing a large quantity of data and counting frequencies, go over the frequency table, looking for empty cells. Each empty cell is assigned a frequency count of 1, and the total count is also incremented by 1. This method pretends that every digram and trigram has been seen at least once. 2. Add 1 to the total count and divide this single 1 among all the empty cells. Each will get a count that’s less than 1 and, as a result, a very small probability. This assigns a very small probability to anything that hasn’t been seen in the training data used for the analysis. An adaptive context-based modeler also maintains tables with the probabilities of all the possible digrams (or trigrams or even longer contexts) of the alphabet and uses the tables to assign a probability to the next symbol S depending on a few symbols immediately preceding it (its context C). The tables are updated all the time as more data is being input, which adapts the probabilities to the particular data being compressed. Such a model is slower and more complex than the static one but produces better compression, since it uses the correct probabilities even when the input has data with probabilities much different from the average. A text that skews letter probabilities is called a lipogram. (Would a computer program without any goto statements be considered a lipogram?) The word comes from the Greek stem .΄... (lipo or leipo) meaning to miss, to lack, combined with the Greek ..΄.΅΅. (gramma), meaning “letter” or “of letters.” Together they form ...o..΄.΅΅..o.. There are not many examples of literary works that are lipograms: 1. Perhaps the best-known lipogram in English is Gadsby, a full-length novel [Wright 39], by Ernest V. Wright, that does not contain any occurrences of the letter E. 2. Alphabetical Africa by Walter Abish (W. W. Norton, 1974) is a readable lipogram where the reader is supposed to discover the unusual writing style while reading. This style has to do with the initial letters of words. The book consists of 52 chapters. In the first, all words begin with a; in the second, words start with either a or b, etc., until, in Chapter 26, all letters are allowed at the start of a word. In the remaining 26 chapters, the letters are taken away one by one. Various readers have commented on how little or how much they have missed the word “the” and how they felt on finally seeing it (in Chapter 20). 3. The novel La Disparition is a 1969 French lipogram by Georges Perec that does not contain the letter E (this letter actually appears several times, outside the main text, in words that the publisher had to include, and these are all printed in red). La Disparition has been translated to English, where it is titled A Void, by Gilbert Adair. Perec also wrote a univocalic (text employing just one vowel) titled Les Revenentes employing only the vowel E. The title of the English translation (by Ian Monk) is The Exeter Text, Jewels, Secrets, Sex. (Perec also wrote a short history of lipograms; see [Motte 98].) 294 5. Statistical Methods A Quotation from the Preface to Gadsby People as a rule will not stop to realize what a task such an attempt actually is. As I wrote along, in long-hand at first, a whole army of little E’s gathered around my desk, all eagerly expecting to be called upon. But gradually as they saw me writing on and on, without even noticing them, they grew uneasy; and, with excited whisperings among themselves, began hopping up and riding on my pen, looking down constantly for a chance to drop off into some word; for all the world like sea birds perched, watching for a passing fish! But when they saw that I had covered 138 pages of typewriter size paper, they slid off unto the floor, walking sadly away, arm in arm; but shouting back: “You certainly must have a hodge-podge of a yarn there without Us! Why, man! We are in every story ever written, hundreds and thousands of times! This is the first time we ever were shut out!” —Ernest V. Wright 4. Gottlob Burmann, a German poet, created our next example of a lipogram. He wrote 130 poems, consisting of about 20,000 words, without the use of the letter R. It is also believed that during the last 17 years of his life, he even omitted this letter from his daily conversation. 5. A Portuguese lipogram is found in five stories written by Alonso Alcala y Herrera, a Portuguese writer, in 1641, each suppressing one vowel. 6. Other examples, in Spanish, are found in the writings of Francisco Navarrete y Ribera (1659), Fernando Jacinto de Zurita y Haro (1654), and Manuel Lorenzo de Lizarazu y Berbinzana (also 1654). An order-N adaptive context-based modeler reads the next symbol S from the input stream and considers the N symbols preceding S the current order-N context C of S. The model then estimates the probability P that S appears in the input data following the particular context C. Theoretically, the larger N, the better the probability estimate (the prediction). To get an intuitive feeling, imagine the case N = 20,000. It is hard to imagine a situation where a group of 20,000 symbols in the input stream is followed by a symbol S, but another group of the same 20,000 symbols, found later in the same input stream, is followed by a different symbol. Thus, N = 20,000 allows the model to predict the next symbol (to estimate its probability) with high accuracy. However, large values of N have three disadvantages: 1. If we encode a symbol based on the 20,000 symbols preceding it, how do we encode the first 20,000 symbols in the input stream? They may have to be written on the output stream as raw ASCII codes, thereby reducing the overall compression. 2. For large values of N, there may be too many possible contexts. If our symbols are the 7-bit ASCII codes, the alphabet size is 27 = 128 symbols. There are therefore 1282 = 16,384 order-2 contexts, 1283 = 2,097,152 order-3 contexts, and so on. The number of contexts grows exponentially, since it is 128N or, in general, AN, where A is the alphabet size. 5.14 PPM 295 I pounded the keys so hard that night that the letter e flew off the part of the machine that hits the paper. Not wanting to waste the night, I went next door to a neighbor who, I knew, had an elaborate workshop in his cellar. He attempted to solder my e back, but when I started to work again, it flew off like a bumblebee. For the rest of the night I inserted each e by hand, and in the morning I took the last dollars from our savings account to buy a new typewriter. Nothing could be allowed to delay the arrival of my greatest triumph. —Sloan Wilson, What Shall We Wear to This Party, (1976)  Exercise 5.34: What is the number of order-2 and order-3 contexts for an alphabet of size 28 = 256?  Exercise 5.35: What would be a practical example of a 16-symbol alphabet? 3. A very long context retains information about the nature of old data. Experience shows that large data files tend to feature topic-specific content. Such a file may contain different distributions of symbols in different parts (a good example is a history book, where one chapter may commonly use words such as “Greek,” “Athens,” and “Troy,” while the following chapter may use “Roman,” “empire,” and “legion”). Such data is termed nonstationary. Better compression can therefore be achieved if the model takes into account the “age” of data i.e, if it assigns less significance to information gathered from old data and more weight to fresh, recent data. Such an effect is achieved by a short context.  Exercise 5.36: Show an example of a common binary file where different parts may have different bit distributions. As a result, relatively short contexts, in the range of 2 to 10, are used in practice. Any practical algorithm requires a carefully designed data structure that provides fast search and easy update, while holding many thousands of symbols and strings (Section 5.14.5). We now turn to the next point in the discussion. Assume a context-based encoder that uses order-3 contexts. Early in the compression process, the word here was seen several times, but the word there is now seen for the first time. Assume that the next symbol is the r of there. The encoder will not find any instances of the order-3 context the followed by r (the r has 0 probability in this context). The encoder may simply write r on the compressed stream as a literal, resulting in no compression, but we know that r was seen several times in the past following the order-2 context he (r has nonzero probability in this context). The PPM method takes advantage of this knowledge. “uvulapalatopharangoplasty” is the name of a surgical procedure to correct sleep apnea. It is rumored to be the longest (English?) word without any e’s. 296 5. Statistical Methods 5.14.1 PPM Principles The central idea of PPM is to use this knowledge. The PPM encoder switches to a shorter context when a longer one has resulted in 0 probability. Thus, PPM starts with an order-N context. It searches its data structure for a previous occurrence of the current context C followed by the next symbol S. If it finds no such occurrence (i.e., if the probability of this particular C followed by this S is 0), it switches to order N - 1 and tries the same thing. Let C be the string consisting of the rightmost N - 1 symbols of C. The PPM encoder searches its data structure for a previous occurrence of the current context C followed by symbol S. PPM therefore tries to use smaller and smaller parts of the context C, which is the reason for its name. The name PPM stands for “prediction with partial string matching.” Here is the process in some detail. The encoder reads the next symbol S from the input stream, looks at the current order-N context C (the last N symbols read), and based on input data that has been seen in the past, determines the probability P that S will appear following the particular context C. The encoder then invokes an adaptive arithmetic coding algorithm to encode symbol S with probability P. In practice, the adaptive arithmetic encoder is a procedure that receives the quantities HighCumFreq[X] and LowCumFreq[X] (Section 5.10) as parameters from the PPM encoder. As an example, suppose that the current order-3 context is the string the, which has already been seen 27 times in the past and was followed by the letters r (11 times), s (9 times), n (6 times), and m (just once). The encoder assigns these cases the probabilities 11/27, 9/27, 6/27, and 1/27, respectively. If the next symbol read is r, it is sent to the arithmetic encoder with a probability of 11/27, and the probabilities are updated to 12/28, 9/28, 6/28, and 1/28. What if the next symbol read is a? The context the was never seen followed by an a, so the probability of this case is 0. This zero-probability problem is solved in PPM by switching to a shorter context. The PPM encoder asks; How many times was the order-2 context he seen in the past and by what symbols was it followed? The answer may be as follows: Seen 54 times, followed by a (26 times), by r (12 times), etc. The PPM encoder now sends the a to the arithmetic encoder with a probability of 26/54. If the next symbol S was never seen before following the order-2 context he, the PPM encoder switches to order-1 context. Was S seen before following the string e? If yes, a nonzero probability is assigned to S depending on how many times it (and other symbols) was seen following e. Otherwise, the PPM encoder switches to order-0 context. It asks itself how many times symbol S was seen in the past, regardless of any contexts. If it was seen 87 times out of 574 symbols read, it is assigned a probability of 87/574. If the symbol S has never been seen before (a common situation at the start of any compression process), the PPM encoder switches to a mode called order -1 context, where S is assigned the fixed probability 1/(size of the alphabet). To predict is one thing. To predict correctly is another. —Unknown 5.14 PPM 297 Table 5.67 shows contexts and frequency counts for orders 4 through 0 after the 11- symbol string xyzzxyxyzzx has been input and encoded. To understand the operation of the PPM encoder, let’s assume that the 12th symbol is z. The order-4 context is now yzzx, which earlier was seen followed by y but never by z. The encoder therefore switches to the order-3 context, which is zzx, but even this hasn’t been seen earlier followed by z. The next lower context, zx, is of order 2, and it also fails. The encoder then switches to order 1, where it checks context x. Symbol x was found three times in the past but was always followed by y. Order 0 is checked next, where z has a frequency count of 4 (out of a total count of 11). Symbol z is therefore sent to the adaptive arithmetic encoder, to be encoded with probability 4/11 (the PPM encoder “predicts” that it will appear 4/11 of the time). Order 4 Order 3 Order 2 Order 1 Order 0 xyzz›x 2 xyz›z 2 xy›z 2 x›y 3 x 4 yzzx›y 1 yzz›x 2 ›x 1 y›z 2 y 3 zzxy›x 1 zzx›y 1 yz›z 2 ›x 1 z 4 zxyx›y 1 zxy›x 1 zz›x 2 z›z 2 xyxy›z 1 xyx›y 1 zx›y 1 ›x 2 yxyz›z 1 yxy›z 1 yx›y 1 (a) Order 4 Order 3 Order 2 Order 1 Order 0 xyzz›x 2 xyz›z 2 xy›z 2 x›y 3 x 4 yzzx›y 1 yzz›x 2 xy›x 1 ›z 1 y 3 ›z 1 zzx›y 1 yz›z 2 y›z 2 z 5 zzxy›x 1 ›z 1 zz›x 2 ›x 1 zxyx›y 1 zxy›x 1 zx›y 1 z›z 2 xyxy›z 1 xyx›y 1 ›z 1 ›x 2 yxyz›z 1 yxy›z 1 yx›y 1 (b) Table 5.67: (a) Contexts and Counts for “xyzzxyxyzzx”. (b) Updated After Another z Is Input. Next, we consider the PPM decoder. There is a fundamental difference between the way the PPM encoder and decoder work. The encoder can always look at the next symbol and base its next step on what that symbol is. The job of the decoder is to find out what the next symbol is. The encoder decides to switch to a shorter context based on what the next symbol is. The decoder cannot mirror this, since it does not know what the next symbol is. The algorithm needs an additional feature that will make it possible for the decoder to stay in lockstep with the encoder. The feature used by PPM is to reserve one symbol of the alphabet as an escape symbol. When the encoder decides to switch to a shorter context, it first writes the escape symbol (arithmetically encoded) on the output stream. The decoder can decode the escape symbol, since it is encoded 298 5. Statistical Methods in the present context. After decoding an escape, the decoder also switches to a shorter context. The worst that can happen with an order-N encoder is to encounter a symbol S for the first time (this happens mostly at the start of the compression process). The symbol hasn’t been seen before in any context, not even in order-0 context (i.e., by itself). In such a case, the encoder ends up sending N +1 consecutive escapes to be arithmetically encoded and output, switching all the way down to order -1, followed by the symbol S encoded with the fixed probability 1/(size of the alphabet). Since the escape symbol may be output many times by the encoder, it is important to assign it a reasonable probability. Initially, the escape probability should be high, but it should drop as more symbols are input and decoded and more information is collected by the modeler about contexts in the particular data being compressed.  Exercise 5.37: The escape is just a symbol of the alphabet, reserved to indicate a context switch. What if the data uses every symbol in the alphabet and none can be reserved? A common example is image compression, where a pixel is represented by a byte (256 grayscales or colors). Since pixels can have any values between 0 and 255, what value can be reserved for the escape symbol in this case? Table 5.68 shows one way of assigning probabilities to the escape symbol (this is variant PPMC of PPM). The table shows the contexts (up to order 2) collected while reading and encoding the 14-symbol string assanissimassa. (In the movie “8 1/2,” Italian children utter this string as a magic spell. They pronounce it assa-neeseemassa.) We assume that the alphabet consists of the 26 letters, the blank space, and the escape symbol, a total of 28 symbols. The probability of a symbol in order -1 is therefore 1/28. Notice that it takes 5 bits to encode 1 of 28 symbols without compression. Each context seen in the past is placed in the table in a separate group together with the escape symbol. The order-2 context as, e.g., was seen twice in the past and was followed by s both times. It is assigned a frequency of 2 and is placed in a group together with the escape symbol, which is assigned frequency 1. The probabilities of as and the escape in this group are therefore 2/3 and 1/3, respectively. Context ss was seen three times, twice followed by a and once by i. These two occurrences are assigned frequencies 2 and 1 and are placed in a group together with the escape, which is now assigned frequency 2 (because it is in a group of 2 members). The probabilities of the three members of this group are therefore 2/5, 1/5, and 2/5, respectively. The justification for this method of assigning escape probabilities is the following: Suppose that context abc was seen ten times in the past and was always followed by x. This suggests that the same context will be followed by the same x in the future, so the encoder will only rarely have to switch down to a lower context. The escape symbol can therefore be assigned the small probability 1/11. However, if every occurrence of context abc in the past was followed by a different symbol (suggesting that the data varies a lot), then there is a good chance that the next occurrence will also be followed by a different symbol, forcing the encoder to switch to a lower context (and thus to emit an escape) more often. The escape is therefore assigned the higher probability 10/20.  Exercise 5.38: Explain the numbers 1/11 and 10/20. 5.14 PPM 299 Order 2 Order 1 Order 0 Context f p Context f p Symbol f p as›s 2 2/3 a› s 2 2/5 a 4 4/19 esc 1 1/3 a› n 1 1/5 s 6 6/19 esc› 2 2/5 n 1 1/19 ss›a 2 2/5 i 2 2/19 ss›i 1 1/5 s› s 3 3/9 m 1 1/19 esc 2 2/5 s› a 2 2/9 esc 5 5/19 s› i 1 1/9 sa›n 1 1/2 esc 3 3/9 esc 1 1/2 n› i 1 1/2 an›i 1 1/2 esc 1 1/2 esc 1 1/2 i› s 1 1/4 ni›s 1 1/2 i› m 1 1/4 esc 1 1/2 esc 2 2/4 is›s 1 1/2 m› a 1 1/2 esc 1 1/2 esc 1 1/2 si›m 1 1/2 esc 1 1/2 im›a 1 1/2 esc 1 1/2 ma›s 1 1/2 esc 1 1/2 Table 5.68: Contexts, Counts (f), and Probabilities (p) for “assanissimassa”. Order 0 consists of the five different symbols asnim seen in the input string, followed by an escape, which is assigned frequency 5. Thus, probabilities range from 4/19 (for a) to 5/19 (for the escape symbol). Wall Street indexes predicted nine out of the last five recessions. —Paul A. Samuelson, Newsweek (19 September 1966) 5.14.2 Examples We are now ready to look at actual examples of new symbols being read and encoded. We assume that the 14-symbol string assanissimassa has been completely input and encoded, so the current order-2 context is “sa”. Here are four typical cases: 1. The next symbol is n. The PPM encoder finds that sa followed by n has been seen before and has probability 1/2. The n is encoded by the arithmetic encoder with this 300 5. Statistical Methods probability, which takes, since arithmetic encoding normally compresses at or close to the entropy, -log2(1/2) = 1 bit. 2. The next symbol is s. The PPM encoder finds that sa was not seen before followed by an s. The encoder therefore sends the escape symbol to the arithmetic encoder, together with the probability (1/2) predicted by the order-2 context of sa. It therefore takes 1 bit to encode this escape. Switching down to order 1, the current context becomes a, and the PPM encoder finds that an a followed by an s was seen before and currently has probability 2/5 assigned. The s is then sent to the arithmetic encoder to be encoded with probability 2/5, which produces another 1.32 bits. In total, 1 + 1.32 = 2.32 bits are generated to encode the s. 3. The next symbol is m. The PPM encoder finds that sa was never seen before followed by an m. It therefore sends the escape symbol to the arithmetic encoder, as in Case 2, generating 1 bit so far. It then switches to order 1, finds that a has never been seen followed by an m, so it sends another escape symbol, this time using the escape probability for the order-1 a, which is 2/5. This is encoded in 1.32 bits. Switching to order 0, the PPM encoder finds m, which has probability 1/19 and sends it to be encoded in -log2(1/19) = 4.25 bits. The total number of bits produced is thus 1+1.32+4.25 = 6.57. 4. The next symbol is d. The PPM encoder switches from order 2 to order 1 to order 0, sending two escapes as in Case 3. Since d hasn’t been seen before, it is not found in order 0, and the PPM encoder switches to order -1 after sending a third escape with the escape probability of order 0, of 5/19 (this produces -log2(5/19) = 1.93 bits). The d itself is sent to the arithmetic encoder with its order -1 probability, which is 1/28, so it gets encoded in 4.8 bits. The total number of bits necessary to encode this first d is 1 + 1.32 + 1.93 + 4.8 = 9.05, more than the five bits that would have been necessary without any compression.  Exercise 5.39: Suppose that Case 4 has actually occurred (i.e., the 15th symbol to be input was a d). Show the new state of the order-0 contexts.  Exercise 5.40: Suppose that Case 4 has actually occurred and the 16th symbol is also a d. How many bits would it take to encode this second d?  Exercise 5.41: Show how the results of the above four cases are affected if we assume an alphabet size of 256 symbols. 5.14.3 Exclusion When switching down from order 2 to order 1, the PPM encoder can use the information found in order 2 in order to exclude certain order-1 cases that are now known to be impossible. This increases the order-1 probabilities and thereby improves compression. The same thing can be done when switching down from any order. Here are two detailed examples. In Case 2, the next symbol is s. The PPM encoder finds that sa was seen before followed by n but not by s. The encoder sends an escape and switches to order 1. The current context becomes a, and the encoder checks to see whether an a followed by an s was seen before. The answer is yes (with frequency 2), but the fact that sa was seen 5.14 PPM 301 before followed by n implies that the current symbol cannot be n (if it were, it would be encoded in order 2). The encoder can therefore exclude the case of an a followed by n in order-1 contexts [we can say that there is no need to reserve “room” (or “space”) for the probability of this case, since it is impossible]. This reduces the total frequency of the order-1 group “a›” from 5 to 4, which increases the probability assigned to s from 2/5 to 2/4. Based on our knowledge from order 2, the s can now be encoded in -log2(2/4) = 1 bit instead of 1.32 (a total of two bits is produced, since the escape also requires 1 bit). Another example is Case 4, modified for exclusions. When switching from order 2 to order 1, the probability of the escape is, as before, 1/2. When in order 1, the case of a followed by n is excluded, increasing the probability of the escape from 2/5 to 2/4. After switching to order 0, both s and n represent impossible cases and can be excluded. This leaves the order 0 with the four symbols a, i, m, and escape, with frequencies 4, 2, 1, and 5, respectively. The total frequency is 12, so the escape is assigned probability 5/12 (1.26 bits) instead of the original 5/19 (1.93 bits). This escape is sent to the arithmetic encoder, and the PPM encoder switches to order -1. Here it excludes all five symbols asnim that have already been seen in order 1 and are therefore impossible in order -1. The d can now be encoded with probability 1/(28-5) . 0.043 (4.52 bits instead of 4.8) or 1/(256 - 5) . 0.004 (7.97 bits instead of 8), depending on the alphabet size. Exact and careful model building should embody constraints that the final answer had in any case to satisfy. —Francis Crick, What Mad Pursuit, (1988) 5.14.4 Four PPM Variants The particular method described earlier for assigning escape probabilities is called PPMC. Four more methods, titled PPMA, PPMB, PPMP, and PPMX, have also been developed in attempts to assign precise escape probabilities in PPM. All five methods have been selected based on the vast experience that the developers had with data compression. The last two are based on Poisson distribution [Witten and Bell 91], which is the reason for the “P” in PPMP (the “X” comes from “approximate,” since PPMX is an approximate variant of PPMP). Suppose that a group of contexts in Table 5.68 has total frequencies n (excluding the escape symbol). PPMA assigns the escape symbol a probability of 1/(n + 1). This is equivalent to always assigning it a count of 1. The other members of the group are still assigned their original probabilities of x/n, and these probabilities add up to 1 (not including the escape probability). PPMB is similar to PPMC with one difference. It assigns a probability to symbol S following context C only after S has been seen twice in context C. This is done by subtracting 1 from the frequency counts. If, for example, context abc was seen three times, twice followed by x and once by y, then x is assigned probability (2-1)/3, and y (which should be assigned probability (1-1)/3 = 0) is not assigned any probability (i.e., does not get included in Table 5.68 or its equivalent). Instead, the escape symbol “gets” the two counts subtracted from x and y, and it ends up being assigned probability 2/3. This method is based on the belief that “seeing twice is believing.” 302 5. Statistical Methods PPMP is based on a different principle. It considers the appearance of each symbol a separate Poisson process. Suppose that there are q different symbols in the input stream. At a certain point during compression, n symbols have been read, and symbol i has been input ci times (so ci = n). Some of the cis are zero (this is the zero-probability problem). PPMP is based on the assumption that symbol i appears according to a Poisson distribution with an expected value (average) .i. The statistical problem considered by PPMP is to estimate q by extrapolating from the n-symbol sample input so far to the entire input stream of N symbols (or, in general, to a larger sample). If we express N in terms of n in the form N = (1+.)n, then a lengthy analysis shows that the number of symbols that haven’t appeared in our n-symbol sample is given by t1.-t2.2+t3.3-· · ·, where t1 is the number of symbols that appeared exactly once in our sample, t2 is the number of symbols that appeared twice, and so on. Hapax legomena: words or forms that occur only once in the writings of a given language; such words are extremely difficult, if not impossible, to translate. In the special case where N is not the entire input stream but the slightly larger sample of size n+1, the expected number of new symbols is t1 1 n -t2 1 n2 +t3 1 n3 -· · ·. This expression becomes the probability that the next symbol is novel, so it is used in PPMP as the escape probability. Notice that when t1 happens to be zero, this expression is normally negative and cannot be used as a probability. Also, the case t1 = n results in an escape probability of 1 and should be avoided. Both cases require corrections to the sum above. PPMX uses the approximate value t1/n (the first term of the sum) as the escape probability. This expression also breaks down when t1 happens to be 0 or n, so in these cases PPMX is modified to PPMXC, which uses the same escape probability as PPMC. Experiments with all five variants show that the differences between them are small. Version X is indistinguishable from P, and both are slightly better than A-B-C. Version C is slightly but consistently better than A and B. It should again be noted that the way escape probabilities are assigned in the A-B-C variants is based on experience and intuition, not on any underlying theory. Experience with these variants indicates that the basic PPM algorithm is robust and is not affected much by the precise way of computing escape probabilities. Variants P and X are based on theory, but even they don’t significantly improve the performance of PPM. 5.14.5 Implementation Details The main problem in any practical implementation of PPM is to maintain a data structure where all contexts (orders 0 through N) of every symbol read from the input stream are stored and can be located fast. The structure described here is a special type of tree, called a trie. This is a tree in which the branching structure at any level is determined by just part of a data item, not by the entire item (page 357). In the case of PPM, an order-N context is a string that includes all the shorter contexts of orders N -1 through 0, so each context effectively adds just one symbol to the trie. Figure 5.69 shows how such a trie is constructed for the string “zxzyzxxyzx” assuming N = 2. A quick glance shows that the tree grows in width but not in depth. Its depth remains N + 1 = 3 regardless of how much input data has been read. Its width grows as more and more symbols are input, but not at a constant rate. Sometimes, no 5.14 PPM 303 new nodes are added, such as in case 10, when the last x is read. At other times, up to three nodes are added, such as in cases 3 and 4, when the second z and the first y are added. Level 1 of the trie (just below the root) contains one node for each symbol read so far. These are the order-1 contexts. Level 2 contains all the order-2 contexts, and so on. Every context can be found by starting at the root and sliding down to one of the leaves. In case 3, for example, the two contexts are xz (symbol z preceded by the order-1 context x) and zxz (symbol z preceded by the order-2 context zx). In case 10, there are seven contexts ranging from xxy and xyz on the left to zxz and zyz on the right. The numbers in the nodes are context counts. The “z,4” on the right branch of case 10 implies that z has been seen four times. The “x,3” and “y,1” below it mean that these four occurrences were followed by x three times and by y once. The circled nodes show the different orders of the context of the last symbol added to the trie. In case 3, for example, the second z has just been read and added to the trie. It was added twice, below the x of the left branch and the x of the right branch (the latter is indicated by the arrow). Also, the count of the original z has been incremented to 2. This shows that the new z follows the two contexts x (of order 1) and zx (order 2). It should now be easy for the reader to follow the ten steps of constructing the tree and to understand intuitively how nodes are added and counts updated. Notice that three nodes (or, in general, N+1 nodes, one at each level of the trie) are involved in each step (except the first few steps when the trie hasn’t reached its final height yet). Some of the three are new nodes added to the trie; the others have their counts incremented. The next point that should be discussed is how the algorithm decides which nodes to update and which to add. To simplify the algorithm, one more pointer is added to each node, pointing backward to the node representing the next shorter context. A pointer that points backward in a tree is called a vine pointer. Figure 5.70 shows the first ten steps in the construction of the PPM trie for the 14-symbol string “assanissimassa”. Each of the ten steps shows the new vine pointers (the dashed lines in the figure) constructed by the trie updating algorithm while that step was executed. Notice that old vine pointers are not deleted; they are just not shown in later diagrams. In general, a vine pointer points from a node X on level n to a node with the same symbol X on level n - 1. All nodes on level 1 point to the root. A node in the PPM trie therefore consists of the following fields: 1. The code (ASCII or other) of the symbol. 2. The count. 3. A down pointer, pointing to the leftmost child of the node. In Figure 5.70, Case 10, for example, the leftmost son of the root is “a,2”. That of “a,2” is “n,1” and that of “s,4” is “a,1”. 4. A right pointer, pointing to the next sibling of the node. The root has no right sibling. The next sibling of node “a,2” is “i,2” and that of “i,2” is “m,1”. 5. A vine pointer. These are shown as dashed arrows in Figure 5.70.  Exercise 5.42: Complete the construction of this trie and show it after all 14 characters have been input. 304 5. Statistical Methods z,1 x,1 z,1 x,1 x,1 z,2 x,1 z,1 z,1 x,1 z2 x,1 z,1 z,1 y,1 y,1 y,1 x,1 z,3 x,1 z,1 z,1 y,1 y,1 z,1 y,1 z,1 x,2 z,3 x,2 z,1 z,1 y,1 y,1 z,1 y,1 x,1 z,1 x,3 z,3 x,2 z,1 z,1 y,1 y,1 z,1 y,1 x,1 z,1 x,1 x,1 x,3 z,3 x,2 z,1 z,1 y,2 y,1 z,1 y,1 x,1 z,1 x,1 x,1 y,1 y,1 x,3 z,4 x,2 z,1 z,1 y,2 y,1 z,2 y,1 x,1 z,1 x,1 x,1 y,1 y,1 z,1 x,4 z,4 x,3 z,1 z,1 y,2 y,1 z,2 y,1 x,2 z,1 x,1 x,1 y,1 y,1 z,1 1. ‘z’ 2. ‘x’ 3. ‘z’ 4. ‘y’ 5. ‘z’ 6. ‘x’ 7. ‘x’ 8. ‘y’ 9. ‘z’ 10. ‘x’ Figure 5.69: Ten Tries of “zxzyzxxyzx”. 5.14 PPM 305 a,1 s,1 s,1 a,1 a,1 s,2 s,1 a,2 s,2 s,1 s,1 s,1 a,1 s,1 s,1 a,1 a,2 s,2 s,1 a,1 s,1 s,1 a,1 n,1 n,1 n,1 a,2 s,2 s,1 s,1 n,1 n,1 i,1 i,1 i,1 a,1 s,1 n,1 a,1 1. ‘a’ 2. ‘s’ 3. ‘s’ 4. ‘a’ 5. ‘n’ 6. ‘i’ Figure 5.70: Part I. First Six Tries of “assanissimassa”. At any step during the trie construction, one pointer, called the base, is maintained that points to the last node added or updated in the previous step. This pointer is shown as a solid arrow in Figure 5.70. Suppose that symbol S has been input and the trie should be updated at this point. The algorithm for adding and/or updating nodes is as follows: 1. Follow the base pointer to node X. Follow the vine pointer from X to Y (notice that Y can be the root). Add S as a new child node of Y and set the base to point to it. However, if Y already has a child node with S, increment the count of that node by 1 (and also set the base to point to it). Call this node A. 2. Repeat the same step but without updating the base. Follow the vine pointer from Y to Z, add S as a new child node of Z, or update an existing child. Call this node B. If there is no vine pointer from A to B, install one. (If both A and B are old nodes, there will already be a vine pointer from A to B.) 3. Repeat until you have added (or incremented) a node at level 1. During these steps, the PPM encoder also collects the counts that are needed to compute the probability of the new symbol S. Figure 5.70 shows the trie after the last 306 5. Statistical Methods a,2 s,3 a,1 s,1 a,1 n,1 n,1 i,1 i,1 s,1 s,1 a,2 s,4 s,1 a,1 s,2 s,1 a,1 n,1 n,1 i,1 i,1 s,1 s,1 s,1 a,2 s,4 s,1 a,1 s,2 s,1 a,1 n,1 n,1 n,1 i,2 i,1 i,1 s,1 s,1 s,1 i,1 i,1 a,2 s,4 a,1 s,2 a,1 n,1 n,1 i,2 i,1 s,1 s,1 s,1 i,1 m,1 i,1 m,1 m,1 7. ‘s’ 8. ‘s’ 9. ‘i’ 10. ‘m’ s,1 s,1 n,1 i,1 s,1 s,1 n,1 i,1 s,1 s,1 n,1 i,1 Figure 5.70: (Continued) Next Four Tries of “assanissimassa”. two symbols s and a were added. In Figure 5.70, Case 13, a vine pointer was followed from node “s,2”, to node “s,3”, which already had the two children “a,1” and “i,1”. The first child was incremented to “a,2”. In Figure 5.70, Case 14, the subtree with the three nodes “s,3”, “a,2”, and “i,1” tells the encoder that a was seen following context ss twice and i was seen following the same context once. Since the tree has two children, the escape symbol gets a count of 2, bringing the total count to 5. The probability of a is therefore 2/5 (compare with Table 5.68). Notice that steps 11 and 12 are not shown. The serious reader should draw the tries for these steps as a voluntary exercise (i.e., without an answer). It is now easy to understand the reason why this particular trie is so useful. Each time a symbol is input, it takes the algorithm at most N +1 steps to update the trie and collect the necessary counts by going from the base pointer toward the root. Adding a symbol to the trie and encoding it takes O(N) steps regardless of the size of the trie. Since N is small (typically 4 or 5), an implementation can be made fast enough for practical use even if the trie is very large. If the user specifies that the algorithm should use exclusions, it becomes more complex, since it has to maintain, at each step, a list of symbols to be excluded. 5.14 PPM 307 a,3 s,6 s,2 a,1 s,3 s,2 a,1 n,1 n,1 n,1 i,2 i,1 i,1 s,1 s,1 s,1 i,1 m,1 i,1 m,1 m,1 13. ‘s’ a,1 a,1 s,1 a,4 s,6 s,2 a,2 s,3 s,2 a,2 n,1 n,1 n,1 i,2 i,1 i,1 s,1 s,1 s,1 i,1 m,1 i,1 m,1 m,1 14. ‘a’ a,1 a,1 s,1 Figure 5.70: (Continued) Final Two Tries of “assanissimassa”. As has been noted, between 0 and 3 nodes are added to the trie for each input symbol encoded (in general, between 0 and N + 1 nodes). The trie can therefore grow very large and fill up any available memory space. One elegant solution, adopted in [Moffat 90], is to discard the trie when it gets full and start constructing a new one. In order to bring the new trie “up to speed” fast, the last 2048 input symbols are always saved in a circular buffer in memory and are used to construct the new trie. This reduces the amount of inefficient code generated when tries are replaced. 5.14.6 PPM* An important feature of the original PPM method is its use of a fixed-length, bounded initial context. The method selects a value N for the context length and always tries to predict (i.e., to assign probability to) the next symbol S by starting with an order- N context C. If S hasn’t been seen so far in context C, PPM switches to a shorter context. Intuitively it seems that a long context (large value of N) may result in better prediction, but Section 5.14 explains the drawbacks of long contexts. In practice, PPM implementations tend to use N values of 5 or 6 (Figure 5.71). The PPM* method, due to [Cleary et al. 95] and [Cleary and Teahan 97], tries to extend the value of N indefinitely. The developers tried to find ways to use unbounded 308 5. Statistical Methods 0 2 Bits per character Maximum context length 3 4 5 6 7 8 9 2 4 6 8 10 12 14 16 Figure 5.71: Compression Ratio as a Function of Maximum Context Length. values for N in order to improve compression. The resulting method requires a new trie data structure and more computational resources than the original PPM, but in return it provides compression improvement of about 6% over PPMC. (In mathematics, when a set S consists of symbols ai, the notation S* is used for the set of all the strings of symbols ai.) One problem with long contexts is the escape symbols. If the encoder inputs the next symbol S, starts with an order-100 context, and does not find any past string of 100 symbols that’s followed by S, then it has to emit an escape and try an order-99 context. Such an algorithm may result in up to 100 consecutive escape symbols being emitted by the encoder, which can cause considerable expansion. It is therefore important to allow for contexts of various lengths, not only very long contexts, and decide on the length of a context depending on the current situation. The only restriction is that the decoder should be able to figure out the length of the context used by the encoder for each symbol encoded. This idea is the main principle behind the design of PPM*. An early idea was to maintain a record, for each context, of its past performance. Such a record can be mirrored by the decoder, so both encoder and decoder can use, at any point, that context that behaved best in the past. This idea did not seem to work and the developers of PPM* were also faced with the task of having to explain why it did not work as expected. The algorithm finally selected for PPM* depends on the concept of a deterministic context. A context is defined as deterministic when it gives only one prediction. For example, the context thisismy is deterministic if every appearance of it so far in the input has been followed by the same symbol. Experiments indicate that if a context C is deterministic, the chance that when it is seen next time, it will be followed by a novel symbol is smaller than what is expected from a uniform prior distribution of the symbols. This feature suggests the use of deterministic contexts for prediction in the new version of PPM. Based on experience with deterministic contexts, the developers have arrived at the following algorithm for PPM*. When the next symbol S is input, search all its contexts 5.14 PPM 309 trying to find deterministic contexts of S. If any such contexts are found, use the shortest of them. If no deterministic contexts are found, use the longest nondeterministic context. The result of this strategy for PPM* is that nondeterministic contexts are used most of the time, and they are almost always 5–6 symbols long, the same as those used by traditional PPM. However, from time to time deterministic contexts are used and they get longer as more input is read and processed. (In experiments performed by the developers, deterministic contexts started at length 10 and became as long as 20– 25 symbols after about 30,000 symbols were input.) The use of deterministic contexts results in very accurate prediction, which is the main contributor to the slightly better performance of PPM* over PPMC. A practical implementation of PPM* has to solve the problem of keeping track of long contexts. Each time a symbol S is input, all its past occurrences have to be checked, together with all the contexts, short and long, deterministic or not, for each occurrence. In principle, this can be done by simply keeping the entire data file in memory and checking back for each symbol. Imagine the symbol in position i in the input file. It is preceded by i - 1 symbols, so i - 1 steps are needed to search and find all its contexts. The total number of steps for n symbols is therefore 1 + 2 + · · · + (n - 1) = n(n - 1)/2. For large n, this amounts to O(n2) complexity—too slow for practical implementations. This problem was solved by a special trie, termed a context-trie, where a leaf node points back to the input string whenever a context is unique. Each node corresponds to a symbol that follows some context and the frequency count of the symbol is stored in the node. PPM* uses the same escape mechanism as the original PPM. The implementation reported in the PPM publications uses the PPMC algorithm to assign probabilities to the various escape symbols. Notice that the original PPM uses escapes less and less over time, as more data is input and more context predictions become available. In contrast, PPM* has to use escapes very often, regardless of the amount of data already input, particularly because of the use of deterministic contexts. This fact makes the problem of computing escape probabilities especially acute. Compressing the entire (concatenated) Calgary corpus by PPM* resulted in an average of 2.34 bpc, compared to 2.48 bpc achieved by PPMC. This represents compression improvement of about 6% because 2.34 is 94.4% of 2.48. 5.14.7 PPMZ The PPMZ variant, originated and implemented by Charles Bloom [Bloom 98], is an attempt to improve the original PPM algorithm. It starts from the premise that PPM is a powerful algorithm that can, in principle, compress data to its entropy, especially when presented with large amounts of input data, but performs less than optimal in practice because of poor handling of features such as deterministic contexts, unbounded-length contexts, and local order estimation. PPMZ attempts to handle these features in an optimal way, and it ends up achieving superior performance. The PPMZ algorithm starts, similar to PPM*, by examining the maximum deterministic context of the current symbol. If no deterministic context is found, the PPMZ encoder executes a local-order-estimation (LOE) procedure, to compute an order in the interval [0, 12] and use it to predict the current symbol as the original PPM algorithm does. In addition, PPMZ uses a secondary model to predict the probabilities of the 310 5. Statistical Methods various escape symbols. The originator of the method noticed that the various PPM implementations compress data to about 2 bpc, where most characters are compressed to 1 bpc each, and the remaining characters represent either the start of the input stream or random data. The natural conclusion is that any small improvements in the probability estimation of the most common characters can lead to significant improvements in the overall performance of the method. We start by discussing the way PPMZ handles unbounded contexts. Figure 5.72a shows a situation where the current character is e and its 12-order context is llassumeth. The context is hashed into a pointer P that points to a linked list. The nodes of the list point to all the 12-character strings in the input stream that happen to hash to the same pointer P. (Each node also has a count field indicating the number of times the string pointed to by the node has been a match.) The encoder follows the pointers, looking for a match whose minimum length varies from context to context. Assuming that the minimum match length in our case is 15, the encoder will find the 15-character match eallassumeth (the preceding w makes this a 16- character match, and it may even be longer). The current character e is encoded with a probability determined by the number of times this match has been found in the past, and the match count (in the corresponding node in the list) is updated. Figure 5.72b shows a situation where no deterministic match is found. The current character is again an e and its order-12 context is the same llassumeth, but the only string llassumeth in the data file is preceded by the three characters ya. The encoder does not find a 15-character match, and it proceeds as follows: (1) It outputs an escape to indicate that no deterministic match has been found. (2) It invokes the LOE procedure to compute an order. (3) It uses the order to predict the current symbol the way ordinary PPM does. (4) It takes steps to ensure that such a case will not happen again. (a) hash xyz ...we_all_assume_then... abc ...we_all_assume_then... (b) hash xyz ...ey_all_assume_then... abc ...we_all_assume_then... Figure 5.72: Unbounded-Length Deterministic Contexts in PPMZ. In the first of these steps, the encoder appends a node to the list and sets it to point to the new 12-character context llassumeth. The second step increments the minimum match length of both contexts by 1 (i.e., to 16 characters). This ensures that these two contexts will be used in the future only when the encoder can match enough characters to distinguish between them. This complex procedure is executed by the PPMZ encoder to guarantee that all the unbounded-length contexts are deterministic. 5.14 PPM 311 Local order estimation is another innovation of PPMZ. Traditional PPM uses the same value for N (the maximum order length, typically 5–6) for all input streams, but a more sophisticated version should attempt to estimate different values of N for different input files or even for different contexts within an input file. The LOE computation performed by PPMZ tries to decide which high-order contexts are unreliable. LOE finds a matching context, examines it in each order, and computes a confidence rating for each order. At first, it seems that the best measure of confidence is the entropy of the context, because the entropy estimates the length of the output in bits. In practice, however, this measure turned out to underestimate the reliability of long contexts. The reason mentioned by the method’s developer is that a certain symbol X may be common in an input stream; yet any specific context may include X only once. The confidence measure finally selected for LOE is based on the probability P of the most probable character in the context. Various formulas involving P were tried, and all resulted in about the same performance. The conclusion was that the best confidence measure for LOE is simply P itself. The last important feature of PPMZ is the way it estimates the escape probabilities. This is called secondary escape estimation or SEE. The main idea is to have an adaptive algorithm where not only the counts of the escapes but also the way the counts are computed are determined by the input stream. In each context, the PPMC method of counting the escapes is first applied. This method counts the number of novel characters (i.e., characters found that were nor predicted) in the context. This information is then used to construct an escape context which, in turn, is used to look up the escape probability in a table. The escape context is a number constructed from the four fields listed here, each quantized to a few bits. Both linear and logarithmic quantization were tried. Linear quantization simply truncates the least-significant bits. Logarithmic quantization computes the logarithm of the number to be quantized. This quantizes large numbers more than small ones, with the result that small values remain distinguishable, while large values may become equal. The four components of the escape context are as follows: 1. The PPM order (which is between 0 and 8), quantized to two bits. 2. The escape count, quantized to two bits. 3. The number of successful matches (total count minus the escape count), quantized to three bits. 4. Ten bits from the two characters xy preceding the current symbol S. Seven bits are taken from x and three bits from y. This number becomes the order-2 escape context. After deleting some bits from it, PPMZ also creates order-1 and order-0 contexts (15 bits and 7 bits long, respectively). The escape contexts are used to update a table of escape counts. Each entry in this table corresponds to matches coded from past PPM contexts that had the same escape contexts. The information in the table is then used in a complex way to construct the escape probability that is sent to the arithmetic coder to code the escape symbol itself. The advantage of this complex method is that it combines the statistics gathered from the long (high order) contexts. These contexts provide high compression but are sparse, causing the original PPM to overestimate their escape probabilities. 312 5. Statistical Methods Applied to the entire (concatenated) Calgary Corpus, PPMZ resulted in an average of 2.119 bpc. This is 10% better than the 2.34 bpc obtained by PPM* and 17% better than the 2.48 bpc achieved by PPMC. 5.14.8 Fast PPM Fast PPM is a PPM variant developed and implemented by [Howard and Vitter 94b] as a compromise between speed and performance of PPM. Even though it is recognized as one of the best (if not the best) statistical compression method, PPM is not very popular because it is slow. Most general-purpose lossless compression software implementations select a dictionary-based method. Fast PPM attempts to isolate those features of PPM that contribute only marginally to compression performance and replace them by approximations. It was hoped that this would speed up execution to make this version competitive with common, commercial lossless compression products. PPM has two main aspects: modeling and encoding. The modeling part searches the contexts of the current symbol to compute its probability. The fast version simplifies that part by eliminating the explicit use of escape symbols, computing approximate probabilities, and simplifying the exclusion mechanism. The encoding part of PPM uses adaptive arithmetic coding. The fast version speeds up this part by using quasiarithmetic coding, a method developed by the same researchers [Howard and Vitter 92c] and not discussed here. The modeling part of fast PPM is illustrated in Table 5.73. We assume that the input stream starts with the string abcbabdbaeabbabe and that the current symbol is the second e (the last character of the string). Part (a) of the table lists the contexts of this character starting with order 3. The order-3 context of this e is bab, which has been seen once in the past, but was followed by a d, so it cannot be used to predict the current e. The encoder therefore skips to order 2, where the context ab was seen three times, but never followed by an e (notice that the d following ab has to be excluded). Skipping to order 1, the encoder finds four different symbols following the order-1 context b. They are c, a, d, and b. Of these, c, d, and b have already been seen, following longer contexts, and are therefore excluded, and the d is designated NF (not found), because we are looking for an e. Skipping to order 0, the encoder finally finds e, following a, b, c, and d, which are all excluded. The point is that both the encoder and decoder of fast PPM can easily generate this table with the information available to them. All that the decoder needs in order to decode the e is the number of NFs (4 in our example) in the table. Part (b) of the table illustrates the situation when the sixth b (there are seven b’s in all) is the current character. It shows that this character can be identified to the decoder by encoding three NFs and writing them on the compressed stream. A different way of looking at this part of fast PPM is to imagine that the encoder generates a list of symbols, starting at the highest order and eliminating duplicates. The list for part (a) of the table consists of dcbae (four NFs followed by an F), while the list for part (b) is cdab (three NFs followed by an F). 5.14 PPM 313 Order Context Symbol Count Action 3 bab d 1 NF, › 2 2 ab c 1 NF d 1 exclude b 1 NF, › 1 1 b c 1 exclude a 3 NF d 1 exclude b 1 exclude, › 0 0 a 5exclude b 7 exclude c 1 exclude d 1 exclude e 1 found Order Context Symbol Count Action 3 eab – – › 2 2 ab c 1 NF d 1 NF, › 1 1 b c 1 exclude a 2 NF d 1 exclude, › 0 0 a 4exclude b 5 found (a) (b) Figure 5.73: Two Examples of Fast PPM For abcbabdbaeabbabe. Temporal reasoning involves both prediction and explanation. Prediction is projection forwards from causes to effects whilst explanation is projection backwards from effects to causes. That is, prediction is reasoning from events to the properties and events they cause, whilst explanation is reasoning from properties and events to events that may have caused them. Although it is clear that a complete framework for temporal reasoning should provide facilities for solving both prediction and explanation problems, prediction has received far more attention in the temporal reasoning literature than explanation. —Murray Shanahan, Proceedings IJCAI 1989 Thus, fast PPM encodes each character by encoding a sequence of NFs, followed by one F (found). It therefore uses a binary arithmetic coder. For increased speed, quasi-arithmetic coding is used, instead of the more common QM coder of Section 5.11. For even faster operation, the quasi-arithmetic coder is used to encode the NFs only for symbols with highest probability, then use a Rice code (Section 10.9) to encode the symbol’s (e or b in our example) position in the remainder of the list. Variants of fast PPM can eliminate the quasi-arithmetic coder altogether (for maximum speed) or use it all the way (for maximum compression). The results of applying fast PPM to the Calgary corpus are reported by the developers and seem to justify its development effort. The compression performance is 2.341 bpc (for the version with just quasi-arithmetic coder) and 2.287 bpc (for the version with both quasi-arithmetic coding and Rice code). This is somewhat worse than the 2.074 bpc achieved by PPMC. However, the speed of fast PPM is about 25,000–30,000 characters per second, compared to about 16,000 characters per second for PPMC—a speedup factor of about 2! 314 5. Statistical Methods 5.15 PAQ PAQ is an open-source high-performance compression algorithm and free software that features sophisticated prediction (modeling) combined with adaptive arithmetic encoding. PAQ encodes individual bits from the input data and its main novelty is its adaptive prediction method, which is based on mixing predictions (or models) obtained from several contexts. PAQ is therefore related to PPM (Section 5.14). The main references are [wikiPAQ 08] and [Mahoney 08] (the latter includes four pdf files with many details about PAQ and its varieties). PAQ has been a surprise to the data compression community. Its algorithm naturally lends itself to all kinds of additions and improvements and this fact, combined with its being open source, inspired many researchers to extend and improve PAQ. In addition to describing the chief features of PAQ, this section discusses the main historical milestones in the rapid development of this algorithm, its versions, subversions, and derivatives. The conclusion at the end of this section is that PAQ may signal a change of paradigm in data compression. In the foreseeable future we may see more extensions of existing algorithms and fewer new, original methods and approaches to compression. The original idea for PAQ is due to Matt Mahoney. He knew that PPM achieves high compression factors and he wondered whether the main principle of PPM, prediction by examining contexts, could be improved by including several predictors (i.e., models of the data) and combining their results in sophisticated ways. As always, there is a price to pay for improved performance, and in the case of PAQ the price is increased execution time and memory requirements. Starting in 2002, Mahoney and others have come up with eight main versions of PAQ, each of which acquired several subversions (or variations), some especially tuned for certain types of data. As of late 2008, the latest version, PAQ8, has about 20 subversions. The PAQ predictor predicts (i.e., computes a probability for) the next input bit by employing several context-based prediction methods (mathematical models of the data). An important feature is that the contexts do not have to be contiguous. Here are the main predictors used by the various versions of PAQ: An order-n context predictor, as used by PPM. The last n bits encoded are examined and the numbers of zeros and 1’s are counted to estimate the probabilities that the next bit will be a 0 or a 1. Whole word order-n context. The context is the latest n whole words input. Nonalphabetic characters are ignored and upper- and lowercase letters are considered identical. This type of prediction makes sense for text files. Sparse contexts. The context consists of certain noncontiguous bytes preceding the current bit. This may be useful in certain types of binary files. A high-order context. The next bit is predicted by examining and counting the bits in the high-order parts of the latest bytes or 16-bit words. This has proved useful for the compression of multimedia files. Two-dimensional context. An image is a rectangle of correlated pixels, so we can expect consecutive rows of an image to be very similar. If a row is c pixels long, then 5.15 PAQ 315 the next pixel to be predicted should be similar to older pixels at distances of c, 2c, and so on. This kind of context also makes sense for tables and spreadsheets. A compressed file generally looks random, but if we know the compression method, we can often see order in the file. Thus, common image compression methods, such as jpeg, tiff, and PNG, produce files with certain structures which can be exploited by specialized predictors. The PAQ encoder examines the beginning of the input file, trying to determine its type (text, jpeg, binary, spreadsheet, etc.). Based on the result, it decides which predictors to use for the file and how to combine the probabilities that they produce for the next input bit. There are three main ways to combine predictions, as follows: 1. In version 1 through 3 of PAQ, each predictor produces a pair of bit counts (n0, n1). The pair for each predictor is multiplied by a weight that depends on the length of the context used by the predictor (the weight for an order-n context is (n+1)2). The weighted pairs are then added and the results normalized to the interval [0, 1], to become meaningful as probabilities. 2. In versions 4 through 6, the weights assigned to the predictors are not fixed but are adjusted adaptively in the direction that would reduce future prediction mistakes. This adjustment tends to favor the more accurate predictors. Specifically, if the predicted bit is b, then weight wi of the ith predictor is adjusted in the following steps ni = n0i + n1i, error = b - p1, wi ‹ wi + (S · n1i - S1 · ni)/(S0 · S1)· error, where n0i and n1i are the counts of zeros and 1’s produced by the predictor, p1 is the probability that the next bit will be a 1, ni = n0i + n1i, and S, S0, and S1 are defined by Equation (5.6). 3. Starting with version 7, each predictor produces a probability rather than a pair of counts. The individual probabilities are then combined with an artificial neural network in the following manner xi = stretch pi1 , p1 = squash(i wixi), where p1 is the probability that the next input bit will be a 1, pi1 is the probability computed by the ith predictor, and the stretch and squash operations are defined as stretch(x) = ln x/(1 - x) and squash(x) = 1/(1 + e-x). Notice that they are the inverse of each other. After each prediction and bit encoding, the ith predictor is updated by adjusting the weight according to wi ‹ wi + ΅xi b - p1 , where ΅ is a constant (typically 0.002 to 0.01) representing the adaptation rate, b is the predicted bit, and b - p1 is the prediction error. In PAQ4 and later versions, the weights are updated by wi ‹ max[0,wi + (b - pi)(S n1i - S1ni)/S0S1], 316 5. Statistical Methods where ni = n0i + n1i. PAQ prediction (or modeling) pays special attention to the difference between stationary and nonstationary data. In the former type, the distribution of symbols (the statistics of the data), remains the same throughout the different parts of the input file. In the latter type, different regions of the input file discuss different topics and consequently feature different symbol distributions. PAQ modeling recognizes that prediction algorithms for nonstationary data must perform poorly for stationary data and vice versa, which is why the PAQ encoder modifies predictions for stationary data to bring them closer to the behavior of nonstationary data. Modeling stationary binary data is straightforward. The predictor initializes the two counts n0 and n1 to zero. If the next bit to be encoded is b, then count nb is incremented by 1. At any point, the probabilities for a 0 and a 1 are p0 = n0/(n0 + n1) and p1 = n1/(n0 + n1). Modeling nonstationary data is more complex and can be done in various ways. The approach considered by PAQ is perhaps the simplest. If the next bit to be encoded is b, then increment count nb by 1 and clear the other counter. Thus, a sequence of consecutive zeros will result in higher and higher values of n0 and in n1 = 0. Essentially, this is the same as predicting that the last input will repeat. Once the next bit of 1 appears in the input, n0 is cleared, n1 is set to 1, and we expect (or predict) a sequence of consecutive 1’s. In general, the input data may be stationary or nonstationary, so PAQ needs to combine the two approaches above; it needs a semi-stationary rule to update the two counters, a rule that will prove itself experimentally over a wide range of input files. The solution? Rather than keep all the counts (as in the stationary model above) or discard all the counts that disagree with the last input (as in the nonstationary model), the semi-stationary model discards about half the counts. Specifically, it keeps at least two counts and discards half of the remaining counts. This modeling has the effect of starting off as a stationary model and becoming nonstationary as more data is input and encoded and the counts grow. The rule is: if the next bit to be encoded is b, then increment nb by 1. If the other counter, n1-b is greater than 2, set it to 1 + n1-b/2 . The modified counts of the ith predictor are denoted by n0i and n1i. The counts from the predictors are combined as follows S0 =  +i win0i, S1 =  +i win1i, (5.6) S = S0 + S1, p0 = S0/S, p1 = S1/S, where wi is the weight assigned to the ith predictor,  is a small positive parameter (whose value is determined experimentally) to make sure that S0 and S1 are never zeros, and the division by S normalizes the counts to probabilities. The predicted probabilities p0 and p1 from the equation above are sent to the arithmetic encoder together with the bit to be encoded. History of PAQ The first version, PAQ1, appeared in early 2002. It was developed and implemented by Matt Mahoney and it employed the following contexts: 5.15 PAQ 317 Eight contexts of length zero to seven bytes. These served as general-purpose models. Recall that PAQ operates on individual bits, not on bytes, so each of these models also includes those bits of the current byte that precede the bit currently being predicted and encoded (there may be between zero and seven such bits). The weight of such a context of length n is (n + 1)2. Two word-oriented contexts of zero or one whole words preceding the currently predicted word. (A word consists of just the 26 letters and is case insensitive.) This constitutes the unigram and bigram models. Two fixed-length record models. These are specialized models for two-dimensional data such as tables and still images. This predictor starts with the byte preceding the current byte and looks for an identical byte in older data. When such a byte is located at a distance of, say, c, the model checks the bytes at distances 2c and 3c. If all four bytes are identical, the model assumes that c is the row length of the table, image, or spreadsheet. One match context. This predictor starts with the eight bytes preceding the current bit. It locates the nearest group of eight (or more) consecutive bytes, examines the bit b that follows this context, and predicts that the current bit will be b. All models except the last one (match) employ the semi-stationary update rule described earlier. PAQ2 was written by Serge Osnach and released in mid 2003. This version adds an SSE (secondary symbol estimation) stage between the predictor and encoder of PAQ1. SSE inputs a short, 10-bit context and the current prediction and outputs a new prediction from a table. The table entry is then updated (in proportion to the prediction error) to reflect the actual bit value. PAQ3 has two subversions, released in late 2003 by Matt Mahoney and Serge Osnach. It includes an improved SSE and three additional sparse models. These employ two-byte contexts that skip over intermediate bytes. PAQ4, again by Matt Mahoney, came out also in late 2003. It used adaptive weighting of 18 contexts, where the weights are updated by wi ‹ max[0,wi + (b - pi)(S n1i - S1ni)/S0S1], where ni = n0i + n1i. PAQ5 (December 2003) and PAQ6 (the same month) represent minor improvements, including a new analog model. With these versions, PAQ became competitive with the best PPM implementations and began to attract the attention of researchers and users. The immediate result was a large number (about 30) of small, incremental improvements and subversions. The main contributors were Berto Destasio, Johan de Bock, Fabio Buffoni, Jason Schmidt, and David A. Scott. These versions added a number of features as follows: A second mixer whose weights were selected by a 4-bit context (consisting of the two most-significant bits of the last two bytes). Six new models designed for audio (8 and 16-bit mono and stereo audio samples), 24-bit color images, and 8-bit data. 318 5. Statistical Methods Nonstationary, run-length models (in addition to the older, semi-stationary model). A special model for executable files with machine code for Intel x86 processors. Many more contexts, including general-purpose, match, record, sparse, analog, and word contexts. In may, 2004, Alexander Ratushnyak threw his hat into the ring and entered the world of PAQ and incremental compression improvements. He named his design PAQAR and started issuing versions in swift succession (seven versions and subversions were issued very quickly). PAQAR extends PAQ by adding many new models, multiple mixers with weights selected by context, an SSE stage to each mixer output, and a preprocessor to improve the compression of Intel executable files (relative CALL and JMP operands are replaced by absolute addresses). The result was significant improvements in compression, at the price of slow execution and larger memory space. Another spinoff of PAQ is PAsQDa, four versions of which were issued by Przemys law Skibi΄nski in early 2005. This algorithm is based on PAQ6 and PAQAR, with the addition of an English dictionary preprocessor (termed a word reducing transform, or WRT, by its developer). WRT replaces English words with a code (up to three bytes long) from a dictionary. It also performs other transforms to model punctuations, capitalizations, and end-of-lines. (It should be noted that several specialized subversions of PAQ8 also employ text preprocessing to achieve minor improvements in the compression of certain text files.) PAsQDa achieved the top ranking on the Calgary corpus but not on most other benchmarks. Here are a few words about one of the terms above. For the purposes of context prediction, the end-of-line (EOL) character is artificial in the sense that it intervenes between symbols that are correlated and spoils their correlation (or similarities). Thus, it makes sense to replace the EOL (which is ASCII character CR, LF, or both) with a space. This idea is due to Malcolm Taylor. Another significant byproduct of PAQ6 is RK (or WinRK), by Malcolm Taylor. This variant of PAQ is based on a mixing algorithm named PWCM (PAQ weighted context mixing) and has achieved first place, surpassing PAQAR and PAsQDa, in many benchmarks. In late 2007, Matt Mahoney released PAQ7, a major rewrite of PAQ6, based on experience gained with PAQAR and PAsQDa. The major improvement is not the compression itself but execution speed, which is up to three times faster than PAQAR. PAQ7 lacks the WRT dictionary and the specialized facilities to deal with Intel x86 executable code, but it does include predictors for jpg, bmp, and tiff image files. As a result, it does not compress executables and text files as well as PAsQDa, but it performs better on data that has text with embedded images, such as MS-excel, MS-Word and PDF. PAQ7 also employs a neural network to combine predictions. PAQ8 was born in early 2006. Currently (early 2009), there are about 20 subversions, many issued in series, that achieve incremental improvements in performance by adding, deleting, and updating many of the features and facilities mentioned above. These subversions were developed and implemented by many of the past players in the PAQ arena, as well as a few new ones. The interested reader is referred to [wikiPAQ 08] for the details of these subversions. Only one series of PAQ subversions, namely the eight algorithms PAQ8HP1 through PAQ8HP8, will be mentioned here. They were re- 5.15 PAQ 319 leased by Alexander Ratushnyak from August, 2006 through January, 2007 specifically to claim improvement steps for the Hutter Prize (Section 5.13). I wrote an experimental PAQ9A in Dec. 2007 based on a chain of 2-input mixers rather than a single mixer as in PAQ8, and an LZP preprocessor for long string matches. The goal was better speed at the cost of some compression, and to allow files larger than 2 GB. It does not compress as well as PAQ8, however, because it has fewer models, just order-n, word, and sparse. Also, the LZP interferes with compression to some extent. I wrote LPAQ1 in July 2007 based on PAQ8 but with fewer models for better speed. (Similar goal to PAQ9A, but a file compressor rather than an archiver). Alexander Ratushnyak wrote several later versions specialized for text which are now #3 on the large text benchmark. —Matt Mahoney to D. Salomon, Dec. 2008 A Paradigm Change Many of the algorithms, techniques, and approaches described in this book have originated in the 1980s and 1990s, which is why these two decades are sometimes referred to as the golden age of data compression. However, the unexpected popularity of PAQ with its many versions, subversions, derivatives, and spinoffs (several are briefly mentioned below) seems to signal a change of paradigm. Instead of new, original algorithms, may we expect more and more small, incremental improvements of existing methods and approaches to compression? No one has a perfect crystal ball, and it is possible that there are clever people out there who have novel ideas for new, complex, and magical compression algorithms and are only waiting for another order of magnitude increase in processor speed and memory size to implement and test their ideas. We’ll have to wait and see. PAQ Derivatives Another reason for the popularity of PAQ is its being free. The algorithm is not encumbered by a patent and existing implementations are free. This has encouraged several researchers and coders to extend PAQ in several directions. The following is a short list of the most notable of these derivatives. WinUDA 0.291, is based on PAQ6 but is faster [UDA 08]. UDA 0.301, is based on the PAQ8I algorithm [UDA 08]. KGB is essentially PAQ6v2 with a GUI (beta version supports PAQ7 compression algorithms). See [KGB 08]. Emilcont, an algorithm based on PAQ6 [Emilcont 08]. Peazip. This is a GUI frontend (for Windows and Linux) for LPAQ and various PAQ8* algorithms [PeaZip 08]. PWCM (PAQ weighted context mixing) is an independently developed closed source implementation of the PAQ algorithm. It is used (as a plugin) in WinRK to win first place in several compression benchmarks. 320 5. Statistical Methods 5.16 Context-Tree Weighting Whatever the input stream is, text, pixels, sound, or anything else, it can be considered a binary string. Ideal compression (i.e., compression at or very near the entropy of the string) would be achieved if we could use the bits that have been input so far in order to predict with certainty (i.e., with probability 1) the value of the next bit. In practice, the best we can hope for is to use history to estimate the probability that the next bit will be 1. The context-tree weighting (CTW) method [Willems et al. 95] starts with a given bit-string bt 1 = b1b2 . . . bt and the d bits that precede it cd = b-d . . . b-2b-1 (the context of bt 1). The two strings cd and bt 1 constitute the input stream. The method uses a simple algorithm to construct a tree of depth d based on the context, where each node corresponds to a substring of cd. The first bit b1 is then input and examined. If it is 1, the tree is updated to include the substrings of cdb1 and is then used to calculate (or estimate) the probability that b1 will be 1 given context cd. If b1 is zero, the tree is updated differently and the algorithm again calculates (or estimates) the probability that b1 will be zero given the same context. Bit b1 and its probability are then sent to an arithmetic encoder, and the process continues with b2. The context bits themselves are written on the compressed stream in raw format. The depth d of the context tree is fixed during the entire compression process, and it should depend on the expected correlation among the input bits. If the bits are expected to be highly correlated, a small d may be enough to get good probability predictions and thus good compression. In thinking of the input as a binary string, it is customary to use the term “source.” We think of the bits of the inputs as coming from a certain information source. The source can be memoryless or it can have memory. In the former case, each bit is independent of its predecessors. In the latter case, each bit depends on some of its predecessors (and, perhaps, also on its successors, but these cannot be used because they are not available to the decoder), so they are correlated. We start by looking at a memoryless source where each bit has probability Pa(1) of being a 1 and probability Pa(0) of being a 0. We set . = Pa(1), so Pa(0) = 1 - . (the subscript a stands for “actual” probability). The probability of a particular string bt 1 being generated by the source is denoted by Pa(bt 1), and it equals the product Pa(bt 1) = t( i=1 Pa(bi). If string bt 1 contains a zeros and b ones, then Pa(bt 1) = (1 - .)a.b. Example: Let t = 5, a = 2, and b = 3. The probability of generating a 5-bit binary string with two zeros and three ones is Pa(b51 ) = (1-.)2.3. Table 5.74 lists the values of Pa(b51 ) for seven values of . from 0 to 1. It is easy to see that the maximum is obtained when . = 3/5. To understand these values intuitively we examine all 32 5-bit numbers. Ten of them consist of two zeros and three ones, so the probability of generating such a string is 10/32 = 0.3125 and the probability of generating a particular string out of these 10 is 0.03125. This number is obtained for . = 1/2. In real-life situations we don’t know the value of ., so we have to estimate it based on what has been input in the past. Assuming that the immediate past string bt 1 consists 5.16 Context-Tree Weighting 321 .: 0 1/5 2/5 1/2 3/5 4/5 5/5 Pa(2, 3): 0 0.00512 0.02304 0.03125 0.03456 0.02048 0 Table 5.74: Seven Values of Pa(2, 3). of a zeros and b ones, it makes sense to estimate the probability of the next bit being 1 by Pe(bt+1 = 1|bt 1) = b a + b , where the subscript e stands for “estimate” (the expression above is read “the estimated probability that the next bit bt+1 will be a 1 given that we have seen string bt 1 is. . . ”). The estimate above is intuitive and cannot handle the case a = b = 0. A better estimate, due to Krichevsky and Trofimov [Krichevsky 81], is called KT and is given by Pe(bt+1 = 1|bt 1) = b + 1/2 a + b + 1. The KT estimate, like the intuitive estimate, predicts a probability of 1/2 for any case where a = b. Unlike the intuitive estimate, however, it also works for the case a = b = 0. The KT Boundary All over the globe is a dark clay-like layer that was deposited around 65 million years ago. This layer is enriched with the rare element iridium. Older fossils found under this KT layer include many dinosaur species. Above this layer (younger fossils) there are no dinosaur species found. This suggests that something that happened around the same time as the KT boundary was formed killed the dinosaurs. Iridium is found in meteorites, so it is possible that a large iridium-enriched meteorite hit the earth, kicking up much iridium dust into the stratosphere. This dust then spread around the earth via air currents and was deposited on the ground very slowly, later forming the KT boundary. This event has been called the “KT Impact” because it marks the end of the Cretaceous Period and the beginning of the Tertiary. The letter “K” is used because “C” represents the Carboniferous Period, which ended 215 million years earlier.  Exercise 5.43: Use the KT estimate to calculate the probability that the next bit will be a zero given string bt 1 as the context. Example: We use the KT estimate to calculate the probability of the 5-bit string 01110. The probability of the first bit being zero is (since there is no context) Pe(0|null) = Pe(0|a=b=0) = 1 - 0 + 1/2 0 + 0 + 1 = 1/2. 322 5. Statistical Methods The probability of the entire string is the product Pe(01110) = Pe(2, 3) = Pe(0|null)Pe(1|0)Pe(1|01)Pe(1|011)Pe(0|0111) = Pe(0|a=b=0)Pe(1|a=1,b=0)Pe(1|a=b=1)Pe(1|a=1,b=2)Pe(0|a=1,b=3) = 1 - 0 + 1/2 0 + 0 + 1 · 0 + 1/2 1 + 0 + 1· 1 + 1/2 1 + 1 + 1· 2 + 1/2 1 + 2 + 1·1 - 3 + 1/2 1 + 3 + 1 = 1 2 · 1 4 · 3 6 · 5 8 · 3 10 = 3 256 . 0.01172. In general, the KT estimated probability of a string with a zeros and b ones is Pe(a, b) = 1/2·3/2 · · · (a - 1/2)·1/2·3/2 · · · (b - 1/2) 1·2·3 · · · (a + b) . (5.7) Table 5.75 lists some values of Pe(a, b) calculated by Equation (5.7). Notice that Pe(a, b) = Pe(b, a), so the table is symmetric. 0 1 2 3 4 5 0 - 1/2 3/8 5/16 35/128 63/256 1 1/2 1/8 1/16 5/128 7/256 21/1024 2 3/8 1/16 3/128 3/256 7/1024 9/2048 3 5/8 5/128 3/256 5/1024 5/2048 45/32768 4 35/128 7/256 7/1024 5/2048 35/32768 35/65536 5 63/256 21/1024 9/2048 45/32768 35/65536 63/262144 Table 5.75: KT Estimates for Some Pe(a, b). Up until now we have assumed a memoryless source. In such a source the probability . that the next bit will be a 1 is fixed. Any binary string, including random ones, is generated by such a source with equal probability. Binary strings that have to be compressed in real situations are generally not random and are generated by a nonmemoryless source. In such a source . is not fixed. It varies from bit to bit, and it depends on the past context of the bit. Since a context is a binary string, all the possible past contexts of a bit can be represented by a binary tree. Since a context can be very long, the tree can include just some of the last bits of the context, to be called the suffix. As an example consider the 42-bit string S = 000101100111010110001101001011110010101100. Let’s assume that we are interested in suffixes of length 3. The first 3 bits of S don’t have long enough suffixes, so they are written raw on the compressed stream. Next we examine the 3-bit suffix of each of the last 39 bits of S and count how many times each suffix is followed by a 1 and how many times by a 0. Suffix 001, for example, is followed 5.16 Context-Tree Weighting 323 twice by a 1 and three times by a 0. Figure 5.76a shows the entire suffix tree of depth 3 for this case (in general, this is not a complete binary tree). The suffixes are read from the leaves to the root, and each leaf is labeled with the probability of that suffix being followed by a 1-bit. Whenever the three most recently read bits are 001, the encoder starts at the root of the tree and follows the edges for 1, 0, and 0. It finds 2/5 at the leaf, so it should predict a probability of 2/5 that the next bit will be a 1, and a probability of 1 - 2/5 that it will be a 0. The encoder then inputs the next bit, examines it, and sends it, with the proper probability, to be arithmetically encoded. Figure 5.76b shows another simple tree of depth 2 that corresponds to the set of suffixes 00, 10, and 1. Each suffix (i.e., each leaf of the tree) is labeled with a probability .. Thus, for example, the probability that a bit of 1 will follow the suffix . . . 10 is 0.3. The tree is the model of the source, and the probabilities are the parameters. In practice, neither the model nor the parameters are known, so the CTW algorithm has to estimate them. 0 0 0 0 0 0 0 1 1 1 1 1 1 1 1/3 2/6 4/7 2/5 2/6 5/6 3/4 2/2 1 0 .1=0.1 1 0 .10=0.3 .11=0.5 (a) (b) Figure 5.76: Two Suffix Trees. Next we get one step closer to real-life situations. We assume that the model is known and the parameters are unknown, and we use the KT estimator to estimate the parameters. As an example we use the model of Figure 5.76b but without the probabilities. We use the string 10|0100110 = 10|b1b2b3b4b5b6b7, where the first two bits are the suffix, to illustrate how the probabilities are estimated with the KT estimator. Bits b1 and b4 have suffix 10, so the probability for leaf 10 of the tree is estimated as the KT probability of substring b1b4 = 00, which is Pe(2, 0) = 3/8 (two zeros and no ones) from Table 5.75. Bits b2 and b5 have suffix 00, so the probability for leaf 00 of the 324 5. Statistical Methods tree is estimated as the KT probability of substring b2b5 = 11, which is Pe(0, 2) = 3/8 (no zeros and two ones) from Table 5.75. Bits b3 = 0, b6 = 1, and b7 = 0 have suffix 1, so the probability for leaf 1 of the tree is estimated as Pe(2, 1) = 1/16 (two zeros and a single one) from Table 5.75. The probability of the entire string 0100110 given the suffix 10 is thus the product 3 8 · 3 8 · 1 16 = 9 1024 . .0088.  Exercise 5.44: Use this example to estimate the probabilities of the five strings 0, 00, 000, 0000, and 00000, assuming that each is preceded by the suffix 00. In the last step we assume that the model, as well as the parameters, are unknown. We construct a binary tree of depth d. The root corresponds to the null context, and each node s corresponds to the substring of bits that were input following context s. Each node thus splits up the string. Figure 5.77a shows an example of a context tree for the string 10|0100110 = 10|b1b2b3b4b5b6b7. 0 0 0 0 0 0 1 1 1 1 1 (a) b7 b7 b7 b3 b6 b3 b6 b3 b6 b1 b2 b3 b1 b4 b1 b4 b1 b4 b2 b5 b2 b5 b2 b5 b4 b5 b6 b7 (b) (4,3) (2,1) (1,0) (1,1) (2,0) (0,2) (1,0) (1,1) (1,0) (1,0) (0,2) (2,2) Pw=7/2048 9/128 3/8 1/2 1/8 1/2 1/2 3/8 1/8 1/16 1/2 5/16 Figure 5.77: (a) A Context Tree. (b) A Weighted Context Tree. 5.16 Context-Tree Weighting 325 Figure 5.77b shows how each node s contains the pair (as, bs), the number of zeros and ones in the string associated with s. The root, for example, is associated with the entire string, so it contains the pair (4, 3). We still have to calculate or estimate a weighted probability Psw for each node s, the probability that should be sent to the arithmetic encoder to encode the string associated with s. This calculation is, in fact, the central part of the CTW algorithm. We start at the leaves because the only thing available in a leaf is the pair (as, bs); there is no suffix. The best assumption that can therefore be made is that the substring consisting of as zeros and bs ones that’s associated with leaf s is memoryless, and the best weighted probability that can be defined for the node is the KT estimated probability Pe(as, bs). We therefore define Psw def = Pe(as, bs) if depth(s) = d. (5.8) Using the weighted probabilities for the leaves, we work our way recursively up the tree and calculate weighted probabilities for the internal nodes. For an internal node s we know a little more than for a leaf, since such a node has one or two children. The children, which are denoted by s0 and s1, have already been assigned weighted probabilities. We consider two cases. If the substring associated with suffix s is memoryless, then Pe(as, bs) is a good weighted probability for it. Otherwise the CTW method claims that the product Ps0 w Ps1 w of the weighted probabilities of the child nodes is a good coding probability (a missing child node is considered, in such a case, to have weighted probability 1). Since we don’t know which of the above cases is true for a given internal node s, the best that we can do is to assign s a weighted probability that’s the average of the two cases above, i.e., Psw def = Pe(as, bs) + Ps0 w Ps1 w 2 if depth(s) < d. (5.9) The last step that needs to be described is the way the context tree is updated when the next bit is input. Suppose that we have already input and encoded the string b1b2 . . . bt-1. Thus, we have already constructed a context tree of depth d for this string, we have used Equations (5.8) and (5.9) to calculate weighted probabilities for the entire tree, and the root of the tree already contains a certain weighted probability. We now input the next bit bt and examine it. Depending on what it is, we need to update the context tree for the string b1b2 . . . bt-1bt. The weighted probability at the root of the new tree will then be sent to the arithmetic encoder, together with bit bt, and will be used to encode bt. If bt = 0, then updating the tree is done by (1) incrementing the as counts for all nodes s, (2) updating the estimated probabilities Pe(as, bs) for all the nodes, and (3) updating the weighted probabilities Pw(as, bs) for all the nodes. If bt = 1, then all the bs should be incremented, followed by updating the estimated and weighted probabilities as above. Figure 5.78a shows how the context tree of Figure 5.77b is updated when bt = 0. 326 5. Statistical Methods (a) (5,3) (2,1) 1 0 (1,0) (1,1) (3,0) (0,2) (1,0) (1,1) (2,0) (1,0) (0,2) (3,2) Pw=.0023345 Pe=.0013732 .0117 .0527 .375 .375 .375 .375 .5 .5 .375 .375 .125 .125 .5 .5 .125 .125 0.0625 0.0625 .5 .5 .3125 .25 (b) (4,4) (2,1) 1 0 (1,0) (1,1) (2,1) (0,2) (1,0) (1,1) (1,1) (1,0) (0,2) (2,3) Pw=.00108337 Pe=.00106811 .0117 .0175 .375 .375 .375 .375 .5 .5 .175 .175 .125 .125 .5 .5 .125 .125 0.0625 0.0625 .5 .5 .0625 .0625 Figure 5.78: Context Trees for bt = 0, 1. Figure 5.78b shows how the context tree of Figure 5.77b is updated when bt = 1.  Exercise 5.45: Construct the context trees with depth 3 for the strings 000|0, 000|00, 000|1, and 000|11. The depth d of the context tree is selected by the user (or is built into both encoder and decoder) and does not change during the compression job. The tree has to be updated for each input bit processed, but this requires updating at most d + 1 nodes. The number of operations needed to process n input bits is thus linear in n. I hoped that the contents of his pockets might help me to form a conclusion. —Arthur Conan Doyle, Memoires of Sherlock Holmes 5.16 Context-Tree Weighting 327 5.16.1 CTW for Text Compression The CTW method has so far been developed for compressing binary strings. In practice, we are normally interested in compressing text, image, and sound streams, and this section discusses one approach to applying CTW to text compression. Each ASCII character consists of seven bits, and all 128 7-bit combinations are used. However, some combinations (such as E and T) are more common than others (such as Z, <, and certain control characters). Also, certain character pairs and triplets (such as TH and THE) appear more often than others. We therefore claim that if bt is a bit in a certain ASCII character X, then the t - 1 bits b1b2 . . . bt-1 preceding it can act as context (even if some of them are not even parts of X but belong to characters preceding X). Experience shows that good results are obtained (1) with contexts of size 12, (2) when seven context trees are used, each to construct a model for one of the seven bits, and (3) if the original KT estimate is modified to the zero-redundancy estimate, defined by Pz e (a, b) def = 1 2Pe(a, b) + 1 4.(a = 0)+ 1 4.(b = 0), where .(true) def = 1 and .(false) def = 0. Another experimental feature is a change in the definition of the weighted probabilities. The original definition, Equation (5.9), is used for the two trees on the ASCII borders (i.e., the ones for bits 1 and 7 of each ASCII code). The weighted probabilities for the five context trees for bits 2–6 are defined by Pw s = Ps0 w Ps1 w . This produces typical compression of 1.8 to 2.3 bits/character on the (concatenated) documents of the Calgary Corpus. Paul Volf [Volf 97] has proposed other approaches to CTW text compression. Statistics are like bikinis. What they reveal is suggestive, but what they conceal is vital. —Aaron Levenstein 6 Dictionary Methods Statistical compression methods use a statistical model of the data, which is why the quality of compression they achieve depends on how good that model is. Dictionarybased compression methods do not use a statistical model, nor do they use variablelength codes. Instead they select strings of symbols and encode each string as a token using a dictionary. The dictionary holds strings of symbols, and it may be static or dynamic (adaptive). The former is permanent, sometimes permitting the addition of strings but no deletions, whereas the latter holds strings previously found in the input stream, allowing for additions and deletions of strings as new input is being read. Given a string of n symbols, a dictionary-based compressor can, in principle, compress it down to nH bits where H is the entropy of the string. Thus, dictionary-based compressors are entropy encoders, but only if the input file is very large. For most files in practical applications, dictionary-based compressors produce results that are good enough to make this type of encoder very popular. Such encoders are also general purpose, performing on images and audio data as well as they perform on text. The simplest example of a static dictionary is a dictionary of the English language used to compress English text. Imagine a dictionary containing perhaps half a million words (without their definitions). A word (a string of symbols terminated by a space or a punctuation mark) is read from the input stream and the dictionary is searched. If a match is found, an index to the dictionary is written into the output stream. Otherwise, the uncompressed word itself is written. (This is an example of logical compression.) As a result, the output stream contains indexes and raw words, and it is important to distinguish between them. One way to achieve this is to reserve an extra bit in every item written. In principle, a 19-bit index is sufficient to specify an item in a 219 = 524,288-word dictionary. Thus, when a match is found, we can write a 20-bit token that consists of a flag bit (perhaps a zero) followed by a 19-bit index. When no match is found, a flag of 1 is written, followed by the size of the unmatched word, followed by the word itself. DOI 10.1007/978-1-84882-903-9_6, © Springer-Verlag London Limited 2010 D. Salomon, G. Motta, Handbook of Data Compression, 5th ed. 330 6. Dictionary Methods Example: Assuming that the word bet is found in dictionary entry 1025, it is encoded as the 20-bit number 0|0000000010000000001. Assuming that the word xet is not found, it is encoded as 1|0000011|01111000|01100101|01110100. This is a 4-byte number where the 7-bit field 0000011 indicates that three more bytes follow. Assuming that the size is written as a 7-bit number, and that an average word size is five characters, an uncompressed word occupies, on average, six bytes (= 48 bits) in the output stream. Compressing 48 bits into 20 is excellent, provided that it happens often enough. Thus, we have to answer the question how many matches are needed in order to have overall compression? We denote the probability of a match (the case where the word is found in the dictionary) by P. After reading and compressing N words, the size of the output stream will be N[20P +48(1-P)] = N[48-28P] bits. The size of the input stream is (assuming five characters per word) 40N bits. Compression is achieved when N[48-28P] < 40N, which implies P >0.29. We need a matching rate of 29% or better to achieve compression.  Exercise 6.1: What compression factor do we get with P = 0.9? As long as the input stream consists of English text, most words will be found in a 500,000-word dictionary. Other types of data, however, may not do that well. A file containing the source code of a computer program may contain “words” such as cout, xor, and malloc that may not be found in an English dictionary. A binary file normally contains gibberish when viewed in ASCII, so very few matches may be found, resulting in considerable expansion instead of compression. This shows that a static dictionary is not a good choice for a general-purpose compressor. It may, however, be a good choice for a special-purpose one. Consider a chain of hardware stores, for example. Their files may contain words such as nut, bolt, and paint many times, but words such as peanut, lightning, and painting will be rare. Special-purpose compression software for such a company may benefit from a small, specialized dictionary containing, perhaps, just a few hundred words. The computers in each branch would have a copy of the dictionary, making it easy to compress files and send them between stores and offices in the chain. In general, an adaptive dictionary-based method is preferable. Such a method can start with an empty dictionary or with a small, default dictionary, add words to it as they are found in the input stream, and delete old words because a big dictionary slows down the search. Such a method consists of a loop where each iteration starts by reading the input stream and breaking it up (parsing it) into words or phrases. It then should search the dictionary for each word and, if a match is found, write a token on the output stream. Otherwise, the uncompressed word should be written and also added to the dictionary. The last step in each iteration checks whether an old word should be deleted from the dictionary. This may sound complicated, but it has two advantages: 1. It involves string search and match operations, rather than numerical computations. Many programmers prefer that. 2. The decoder is simple (this is an asymmetric compression method). In statistical compression methods, the decoder is normally the exact opposite of the encoder (symmetric compression). In an adaptive dictionary-based method, however, the decoder has to read its input stream, determine whether the current item is a token or uncompressed data, use tokens to obtain data from the dictionary, and output the final, uncompressed 6.1 String Compression 331 data. It does not have to parse the input stream in a complex way, and it does not have to search the dictionary to find matches. Many programmers like that, too. J. Ziv A. Lempel T. Welch The LZW Trio. Having one’s name attached to a scientific discovery, technique, or phenomenon is considered a special honor in science. Having one’s name associated with an entire field of science is even more so. This is what happened to Jacob Ziv and Abraham Lempel. In the late 1970s these two researchers developed the first methods, LZ77 and LZ78, for dictionary-based compression. Their ideas have been a source of inspiration to many researchers who generalized, improved, and combined them with RLE and statistical methods to form many popular lossless compression methods for text, images, and audio. Close to 30 such methods are described in this chapter, some in great detail. Of special interest is the popular LZW algorithm, partly devised by Terry Welch (Section 6.13), which has extended LZ78 and made it practical and popular. I love the dictionary, Kenny, it’s the only book with the words in the right place. —Paul Reynolds as Colin Mathews in Press Gang (1989) 6.1 String Compression In general, compression methods based on strings of symbols can be more efficient than methods that compress individual symbols. To understand this, the reader should first review Exercise A.4. This exercise shows that in principle, better compression is possible if the symbols of the alphabet have very different probabilities of occurrence. We use a simple example to show that the probabilities of strings of symbols vary more than the probabilities of the individual symbols constituting the strings. We start with a 2-symbol alphabet a1 and a2, with probabilities P1 = 0.8 and P2 = 0.2, respectively. The average probability is 0.5, and we can get an idea of the variance (how much the individual probabilities deviate from the average) by calculating the sum of absolute differences |0.8 - 0.5| + |0.2 - 0.5| = 0.6. Any variable-length code would assign 1-bit codes to the two symbols, so the average length of the code is one bit per symbol. We now generate all the strings of two symbols. There are four of them, shown in Table 6.1a, together with their probabilities and a set of Huffman codes. The average probability is 0.25, so a sum of absolute differences similar to the one above yields |0.64 - 0.25| + |0.16 - 0.25| + |0.16 - 0.25| + |0.04 - 0.25| = 0.78. 332 6. Dictionary Methods String Probability Code a1a1 0.8 Χ 0.8 = 0.64 0 a1a2 0.8 Χ 0.2 = 0.16 11 a2a1 0.2 Χ 0.8 = 0.16 100 a2a2 0.2 Χ 0.2 = 0.04 101 (a) Str. Variance Avg. size size of prob. of code 1 0.6 1 2 0.78 0.78 3 0.792 0.728 (c) String Probability Code a1a1a1 0.8 Χ 0.8 Χ 0.8 = 0.512 0 a1a1a2 0.8 Χ 0.8 Χ 0.2 = 0.128 100 a1a2a1 0.8 Χ 0.2 Χ 0.8 = 0.128 101 a1a2a2 0.8 Χ 0.2 Χ 0.2 = 0.032 11100 a2a1a1 0.2 Χ 0.8 Χ 0.8 = 0.128 110 a2a1a2 0.2 Χ 0.8 Χ 0.2 = 0.032 11101 a2a2a1 0.2 Χ 0.2 Χ 0.8 = 0.032 11110 a2a2a2 0.2 Χ 0.2 Χ 0.2 = 0.008 11111 (b) Table 6.1: Probabilities and Huffman Codes for a Two-Symbol Alphabet. The average size of the Huffman code is 1Χ0.64+2Χ0.16+3Χ0.16+3Χ0.04 = 1.56 bits per string, which is 0.78 bits per symbol. In the next step we similarly create all eight strings of three symbols. They are shown in Table 6.1b, together with their probabilities and a set of Huffman codes. The average probability is 0.125, so a sum of absolute differences similar to the ones above yields |0.512 - 0.125| + 3|0.128 - 0.125| + 3|0.032 - 0.125| + |0.008 - 0.125| = 0.792. The average size of the Huffman code in this case is 1 Χ 0.512 + 3 Χ 3 Χ 0.128 + 3 Χ 5 Χ 0.032 + 5 Χ 0.008 = 2.184 bits per string, which equals 0.728 bits per symbol. As we keep generating longer and longer strings, the probabilities of the strings differ more and more from their average, and the average code size gets better (Table 6.1c). This is why a compression method that compresses strings, rather than individual symbols, can, in principle, yield better results. This is also the reason why the various dictionary-based methods are in general better and more popular than the Huffman method and its variants (see also Section 7.19). The above conclusion is a fundamental result of rate-distortion theory, that part of information theory that deals with data compression. 6.2 Simple Dictionary Compression 333 6.2 Simple Dictionary Compression The topic of this section is a simple, two-pass method, related to me by Ismail Mohamed (see Preface to the 3rd edition of the complete reference). The first pass reads the source file and prepares a list of all the different bytes found. The second pass uses this list to actually compress the data bytes. Here are the steps in detail. 1. The source file is read and a list is prepared of the distinct bytes encountered. For each byte, the number of times it occurs in the source file (its frequency) is also included in the list. 2. The list is sorted in descending order of the frequencies. Thus, it starts with byte values that are common in the file, and it ends with bytes that are rare. Since the list consists of distinct bytes, it can have at most 256 elements. 3. The sorted list becomes the dictionary. It is written on the compressed file, preceded by its length (a 1-byte integer). 4. The source file is read again byte by byte. Each byte is located in the dictionary (by a direct search) and its index is noted. The index is a number in the interval [0, 255], so it requires between 1 and 8 bits (but notice that most indexes will normally be small numbers because common byte values are stored early in the dictionary). The index is written on the compressed file, preceded by a 3-bit code denoting the index’s length. Thus, code 000 denotes a 1-bit index, code 001 denotes a 2-bit index, and so on up to code 111, which denotes an 8-bit index. The compressor maintains a short, 2-byte buffer where it collects the bits to be written on the compressed file. When the first byte of the buffer is filled, it is written on the file and the second byte is moved to the first byte. Decompression is straightforward. The decompressor starts by reading the length of the dictionary, then the dictionary itself. It then decodes each byte by reading its 3-bit code, followed by its index value. The index is used to locate the next data byte in the dictionary. Compression is achieved because the dictionary is sorted by the frequency of the bytes. Each byte is replaced by a quantity of between 4 and 11 bits (a 3-bit code followed by 1 to 8 bits). A 4-bit quantity corresponds to a compression ratio of 0.5, while an 11-bit quantity corresponds to a compression ratio of 1.375, an expansion. The worst case is a file where all 256 byte values occur and have a uniform distribution. The compression ratio in such a case is the average (2 Χ 4 + 2 Χ 5 + 4 Χ 6 + 8 Χ 7 + 16 Χ 8 + 32 Χ 9 + 64 Χ 10 + 128 Χ 11)/(256 Χ 8) = 2562/2048 = 1.2509765625, indicating expansion! (Actually, slightly worse than 1.25, because the compressed file also includes the dictionary, whose length in this case is 257 bytes.) Experience indicates typical compression ratios of about 0.5. The probabilities used here were obtained by counting the numbers of codes of various sizes. Thus, there are two 4-bit codes 000|0 and 000|1, two 5-bit codes 001|10 and 001|11, four 6-bit codes 010|100, 010|101, 010|110 and 010|111, eight 7-bit codes 011|1000, 011|1001, 011|1010, 011|1011, 011|1100, 011|1101, 011|1110, and 011|1111, and so on, up to 128 11-bit codes. 334 6. Dictionary Methods The downside of the method is slow compression; a result of the two passes the compressor performs combined with the slow search (slow, because the dictionary is not sorted by byte values, so binary search cannot be used). Decompression, in contrast, is not slow.  Exercise 6.2: Design a reasonable organization for the list maintained by this method. 6.3 LZ77 (Sliding Window) An important source of redundancy in digital data is repeating phrases. This type of redundancy exists in all types of data—audio, still images, and video—but is easiest to visualize for text. Figure 6.2 lists two paragraphs from this book and makes it obvious that certain phrases, such as words, parts of words, and other bits and pieces of text repeat several times or even many times. Phrases tend to repeat in the same paragraph, but repetitions are often found in paragraphs that are distant from one another. Thus, phrases that occur again and again in the data constitute the basis of all the dictionarybased compression algorithms. Statistical compression methods use a statistical model of the data, which is why the quality of compression they achieve depends on how good that model is. Dictionarybased compression methods do not use a statistical model, nor do they use variable-size codes. Instead they select strings of symbols and encode each string as a token using a dictionary. The dictionary holds strings of symbols, and it may be static or dynamic (adaptive). The former is permanent, sometimes allowing the addition of strings but no deletions, whereas the latter holds strings previously found in the input stream, allowing for additions and deletions of strings as new input is being read. It is generally agreed that an invention or a process is patentable but a mathematical concept, calculation, or proof is not. An algorithm seems to be an abstract mathematical concept that should not be patentable. However, once the algorithm is implemented in software (or in firmware) it may not be possible to separate the algorithm from its implementation. Once the implementation is used in a new product (i.e., an invention), that product—including the implementation (software or firmware) and the algorithm behind it—may be patentable. [Zalta 88] is a general discussion of algorithm patentability. Several common data compression algorithms, most notably LZW, have been patented; and the LZW patent is discussed here in some detail. Figure 6.2: Repeating Phrases in Text. The principle of LZ77 (sometimes also referred to as LZ1) [Ziv and Lempel 77] is to use part of the previously-seen input stream as the dictionary. The encoder maintains a window to the input stream and shifts the input in that window from right to left as strings of symbols are being encoded. Thus, the method is based on a sliding window. The window below is divided into two parts. The part on the left is the search buffer. This is the current dictionary, and it includes symbols that have recently been input and encoded. The part on the right is the look-ahead buffer, containing text yet to be 6.3 LZ77 (Sliding Window) 335 encoded. In practical implementations the search buffer is some thousands of bytes long, while the look-ahead buffer is only tens of bytes long. The vertical bar between the t and the e below represents the current dividing line between the two buffers. We assume that the text sirsideastmaneasilyt has already been compressed, while the text easesseasickseals still needs to be compressed. ‹ coded text. . . sirsideastmaneasilyt|easesseasickseals. . .‹ text to be read The encoder scans the search buffer backwards (from right to left) looking for a match for the first symbol e in the look-ahead buffer. It finds one at the e of the word easily. This e is at a distance (offset) of 8 from the end of the search buffer. The encoder then matches as many symbols following the two e’s as possible. Three symbols eas match in this case, so the length of the match is 3. The encoder then continues the backward scan, trying to find longer matches. In our case, there is one more match, at the word eastman, with offset 16, and it has the same length. The encoder selects the longest match or, if they are all the same length, the last one found, and prepares the token (16, 3, e). Selecting the last match, rather than the first one, simplifies the encoder, because it only has to keep track of the latest match found. It is interesting to note that selecting the first match, while making the program somewhat more complex, also has an advantage. It selects the smallest offset. It would seem that this is not an advantage, because a token should have room for the largest possible offset. However, it is possible to follow LZ77 with Huffman, or some other statistical coding of the tokens, where small offsets are assigned shorter codes. This method, proposed by Bernd Herd, is called LZH. Having many small offsets implies better compression in LZH.  Exercise 6.3: How does the decoder know whether the encoder selects the first match or the last match? In general, an LZ77 token has three parts: offset, length, and next symbol in the look-ahead buffer (which, in our case, is the second e of the word teases). The offset can also be thought of as the distance between the previous and the current occurrences of the string being compressed. This token is written on the output stream, and the window is shifted to the right (or, alternatively, the input stream is moved to the left) four positions: three positions for the matched string and one position for the next symbol. ...sirsideastmaneasilytease|sseasickseals....... If the backward search yields no match, an LZ77 token with zero offset and length and with the unmatched symbol is written. This is also the reason a token has a third component. Tokens with zero offset and length are common at the beginning of any compression job, when the search buffer is empty or almost empty. The first five steps in encoding our example are the following: |sirsideastman . (0,0,“s”) s|irsideastmane . (0,0,“i”) si|rsideastmanea . (0,0,“r”) sir|sideastmaneas . (0,0,“”) sir|sideastmaneasi . (4,2,“d”) 336 6. Dictionary Methods  Exercise 6.4: What are the next two steps? Clearly, a token of the form (0, 0, . . .), which encodes a single symbol, does not provide good compression. It is easy to estimate its length. The size of the offset is log2 S , where S is the length of the search buffer. In practice, the search buffer may be a few thousand bytes long, so the offset size is typically 10–12 bits. The size of the “length” field is similarly log2(L - 1) , where L is the length of the look-ahead buffer (see below for the -1). In practice, the look-ahead buffer is only a few tens of bytes long, so the size of the “length” field is just a few bits. The size of the “symbol” field is typically 8 bits, but in general, it is log2 A , where A is the alphabet size. The total size of the 1-symbol token (0, 0, . . .) may typically be 11 + 5 + 8 = 24 bits, much longer than the raw 8-bit size of the (single) symbol it encodes. Here is an example showing why the “length” field may be longer than the size of the look-ahead buffer: ...Mr.alfeastmaneasilygrowsalf|alfainhisgarden... . The first symbol a in the look-ahead buffer matches the five a’s in the search buffer. It seems that the two extreme a’s match with a length of 3 and the encoder should select the last (leftmost) of them and create the token (28,3,“a”). In fact, it creates the token (3,4,“”). The four-symbol string alfa in the look-ahead buffer is matched with the last three symbols alf in the search buffer and the first symbol a in the look-ahead buffer. The reason for this is that the decoder can handle such a token naturally, without any modifications. It starts at position 3 of its search buffer and copies the next four symbols, one by one, extending its buffer to the right. The first three symbols are copies of the old buffer contents, and the fourth one is a copy of the first of those three. The next example is even more convincing (and only somewhat contrived): ...alfeastmaneasilyyellsA|AAAAAAAAAAAAAAAH... . The encoder creates the token (1,9,A), matching the first nine copies of A in the lookahead buffer and including the tenth A. This is why, in principle, the length of a match can be up to the size of the look-ahead buffer minus 1. The decoder is much simpler than the encoder (LZ77 is therefore an asymmetric compression method). It has to maintain a buffer, equal in size to the encoder’s window. The decoder inputs a token, finds the match in its buffer, writes the match and the third token field on the output stream, and shifts the matched string and the third field into the buffer. This implies that LZ77, or any of its variants, is useful in cases where a file is compressed once (or just a few times) and is decompressed often. A rarely-used archive of compressed files is a good example. At first it seems that this method does not make any assumptions about the input data. Specifically, it does not pay attention to any symbol frequencies. A little thinking, however, shows that because of the nature of the sliding window, the LZ77 method always compares the look-ahead buffer to the recently-input text in the search buffer and never to text that was input long ago (and has therefore been flushed out of the search buffer). Thus, the method implicitly assumes that patterns in the input data occur close together. Data that satisfies this assumption will compress well. The basic LZ77 method was improved in several ways by researchers and programmers during the 1980s and 1990s. One way to improve it is to use variable-size “offset” 6.3 LZ77 (Sliding Window) 337 and “length” fields in the tokens. Another way is to increase the sizes of both buffers. Increasing the size of the search buffer makes it possible to find better matches, but the trade-off is an increased search time. A large search buffer therefore requires a more sophisticated data structure that allows for fast search (Section 6.13.2). A third improvement has to do with sliding the window. The simplest approach is to move all the text in the window to the left after each match. A faster method is to replace the linear window with a circular queue, where sliding the window is done by resetting two pointers (Section 6.3.1). Yet another improvement is adding an extra bit (a flag) to each token, thereby eliminating the third field (Section 6.4). Of special notice is the hash table employed by the Deflate algorithm (Section 6.25.3) to search for matches. 6.3.1 A Circular Queue The circular queue is a basic data structure. Physically, it is a linear array, but it is used as a circular array. Figure 6.3 illustrates a simple example. It shows a 16-byte array with characters appended at the “end” and deleted from the “start.” Both the start and end positions move, and two pointers, s and e, point to them all the time. In (a) the queue consists of the eight characters sideast, with the rest of the buffer empty. In (b) all 16 bytes are occupied, and e points to the end of the buffer. In (c), the first letter s has been deleted and the l of easily inserted. Notice how pointer e is now located to the left of s. In (d), the two letters id have been deleted just by moving the s pointer; the characters themselves are still present in the array but have been effectively deleted. In (e), the two characters y have been appended and the e pointer moved. In (f), the pointers show that the buffer ends at teas and starts at tman. Inserting new characters into the circular queue and moving the pointers is thus equivalent to shifting the contents of the queue. No actual shifting or moving is necessary, though. sideast ^s ^e (a) sideastmaneasi ^s ^e (b) lideastmaneasi ^e^s (c) lideastmaneasi ^e ^s (d) lyeastmaneasi ^e^s (e) lyteastmaneasi ^e^s (f) Figure 6.3: A Circular Queue. More information on circular queues can be found in most texts on data structures. From the dictionary Circular. (1) Shaped like or nearly like a circle. (2) Defining one word in terms of another that is itself defined in terms of the first word. (3) Addressed or distributed to a large number of persons. 338 6. Dictionary Methods 6.3.2 LZR The data compression literature is full of references to LZ77 and LZ78. There is, however, also an LZ76 algorithm [Lempel and Ziv 76]. It is little known because it deals with the complexity of text, rather than with compressing the text. The point is that LZ76 measures complexity by looking for previously-found strings of text. It can therefore be extended to become a text compression algorithm. The LZR method described here, proposed by [Rodeh et al. 81], is such an extension. LZR is a variant of the basic LZ77 method, where the lengths of both the search and look-ahead buffers are unbounded. In principle, such a method will always find the best possible match to the string in the look-ahead buffer, but it is immediately obvious that any practical implementation of LZR will have to deal with the finite memory space and run-time available. The space problem can be handled by allocating more and more space to the buffers until no more space is left, and then either (1) keeping the buffers at their current sizes or (2) clearing the buffers and starting with empty buffers. The time constraint may be more severe. The basic algorithm of LZ76 uses linear search and requires O(n2) time to process an input string of n symbols. The developers of LZR propose to reduce these space and time complexities to O(n) by employing the special suffix trees proposed by [McCreight 76], but readers of this chapter will have many opportunities to see that other LZ methods, such as LZ78 and its varieties, offer similar compression performance, while being simpler to implement. The data structure developed by McCreight is based on a complicated multiway tree. Those familiar with tree data structures know that the basic operations on a tree are (1) adding a new node, (2) modifying an existing node, and (3) deleting a node. Of these, deletion is normally the most problematic since it tends to change the entire structure of a tree. Thus, instead of deleting nodes when the symbols represented by them overflow off the left end of the search buffer, LZR simply marks a node as deleted. It also constructs several trees and deletes a tree when all its nodes have beed marked as deleted (i.e., when all the symbols represented by the tree have slid off the search buffer). As a result, LZR implementation is complex. Another downside of LZR is the sizes of the offset and length fields of an output triplet. With buffers of unlimited sizes, these fields may become big. LZR handles this problem by encoding these fields in a variable-length code. Specifically, the Even–Rodeh codes of Section 3.7 are used to encode these fields. The length of the Even–Rodeh code of the integer n is close to 2 log2 n, so even for n values in the millions, the codes are about 20 bits long. 6.4 LZSS 339 6.4 LZSS LZSS is an efficient variant of LZ77 developed by Storer and Szymanski in 1982 [Storer and Szymanski 82]. It improves LZ77 in three directions: (1) It holds the look-ahead buffer in a circular queue, (2) it creates tokens with two fields instead of three, and (3) it holds the search buffer (the dictionary) in a binary search tree, (this data structure was incorporated in LZSS by [Bell 86]). The second improvement of LZSS over LZ77 is in the tokens created by the encoder. An LZSS token contains just an offset and a length. If no match was found, the encoder emits the uncompressed code of the next symbol instead of the wasteful three-field token (0, 0, . . .). To distinguish between tokens and uncompressed codes, each is preceded by a single bit (a flag). In practice, the search buffer may be a few thousand bytes long, so the offset field would typically be 11–13 bits. The size of the look-ahead buffer should be selected such that the total size of a token would be 16 bits (2 bytes). For example, if the search buffer size is 2 Kbyte (= 211), then the look-ahead buffer should be 32 bytes long (= 25). The offset field would be 11 bits long and the length field, 5 bits (the size of the look-ahead buffer). With this choice of buffer sizes the encoder will emit either 2-byte tokens or 1-byte uncompressed ASCII codes. But what about the flag bits? A good practical idea is to collect eight output items (tokens and ASCII codes) in a small buffer, then output one byte consisting of the eight flags, followed by the eight items (which are 1 or 2 bytes long each). The third improvement has to do with binary search trees. A binary search tree is a binary tree where the left subtree of every node A contains nodes smaller than A, and the right subtree contains nodes greater than A. Since the nodes of our binary search trees contain strings, we first need to know how to compare two strings and decide which one is “bigger.” This is easily understood by imagining that the strings appear in a dictionary or a lexicon, where they are sorted alphabetically. Clearly, the string rote precedes the string said since r precedes s (even though o follows a), so we consider rote smaller than said. This concept is called lexicographic order (ordering strings lexicographically). What about the string abc? Most modern computers use ASCII codes to represent characters (although more and more use Unicode, discussed in Section 11.12, and some older IBM, Amdahl, Fujitsu, and Siemens mainframe computers use the old, 8-bit EBCDIC code developed by IBM), and in ASCII the code of a blank space precedes those of the letters, so a string that starts with a space will be smaller than any string that starts with a letter. In general, the collating sequence of the computer determines the sequence of characters arranged from small to big. Figure 6.4 shows two examples of binary search trees. Notice the difference between the (almost) balanced tree in Figure 6.4a and the skewed one in Figure 6.4b. They contain the same 14 nodes, but they look and behave very differently. In the balanced tree any node can be found in at most four steps. In the skewed tree up to 14 steps may be needed. In either case, the maximum number of steps needed to locate a node equals the height of the tree. For a skewed tree (which is really the same as a linked list), the height is the number of elements n; for a balanced tree, the height is log2 n , a much smaller number. More information on the properties 340 6. Dictionary Methods mine slam obey come went slow fast hers left from plug stay brim step mine obey slam come went slow fast hers left from plug brim stay step (a) (b) Figure 6.4: Two Binary Search Trees. of binary search trees may be found in any text on data structures. Here is an example showing how a binary search tree can be used to speed up the search of the dictionary. We assume an input stream with the short sentence sideastmanclumsilyteasesseasickseals. To keep the example simple, we assume a window of a 16-byte search buffer followed by a 5-byte look-ahead buffer. After the first 16 + 5 characters have been input, the sliding window is sideastmanclum|silyteasesseasickseals with the string teasesseasickseals still waiting to be input. The encoder scans the search buffer, creating the 12 five-character strings of Table 6.5 (12 since 16 - 5 + 1 = 12), which are inserted into the binary search tree, each with its offset. The first symbol in the look-ahead buffer is s, so the encoder searches the tree for strings that start with an s. Two are found, at offsets 16 and 10, and the first of them, side (at offset 16) provides a longer match. (We now have to sidetrack and discuss the case where a string in the tree completely matches that in the look-ahead buffer. In that case the encoder should go back to the search buffer, to attempt to match longer strings. In principle, the maximum length of a match can be L - 1.) In our example, the match is of length 2, and the 2-field token (16, 2) is emitted. The encoder now has to slide the window two positions to the right, and update the tree. The new window is 6.4 LZSS 341 side 16 idea 15 deas 14 east 13 eastm 12 astma 11 stman 10 tman 09 manc 08 ancl 07 nclu 06 clum 05 Table 6.5: Five-Character Strings. sideastmanclumsi|lyteasesseasickseals The tree should be updated by deleting strings side and idea, and inserting the new strings clums and lumsi. If a longer, k-letter, string is matched, the window has to be shifted k positions, and the tree should be updated by deleting k strings and adding k new strings, but which ones? A little thinking shows that the k strings to be deleted are the first ones in the search buffer before the shift, and the k strings to be added are the last ones in it after the shift. A simple procedure for updating the tree is to prepare a string consisting of the first five letters in the search buffer, find it in the tree, and delete it. Then slide the buffer one position to the right (or shift the data to the left), prepare a string consisting of the last five letters in the search buffer, and append it to the tree. This should be repeated k times. Since each update deletes and adds the same number of strings, the tree size never changes, except at the start and end of the compression. It always contains T nodes, where T is the length of the search buffer minus the length of the look-ahead buffer plus 1 (T = S - L + 1). The shape of the tree, however, may change significantly. As nodes are being added and deleted, the tree may change its shape between a completely skewed tree (the worst case for searching) and a balanced one, the ideal shape for searching. When the encoder starts, it outputs the first L symbols of the input in raw format and shifts them from the look-ahead buffer to the search buffer. It then fills up the remainder of the search buffer with blanks. In our example, the initial tree will consist of the five strings s, si, sid, sid, and side. As compression progresses, the tree grows to S - L + 1 = 12 nodes, and stays at that size until it shrinks toward the end. 6.4.1 LZB LZB is an extension of LZSS. This method, proposed in [Bell 87], is the result of evaluating and comparing several data structures and variable-length codes with an eye toward improving the performance of LZSS. In LZSS, the length of the offset field is fixed and is large enough to accommodate any pointers to the search buffer. If the size of the search buffer is S, then the offset 342 6. Dictionary Methods field is s def = log2 S bits long. However, when the encoder starts, the search buffer is empty, so the offset fields of the first tokens output do not use all the bits allocated to them. In principle, the size of the offset field should start at one bit. This is enough as long as there are only two symbols in the search buffer. The offset should then be increased to two bits until the search buffer contains four symbols, and so on. As more and more symbols are shifted into the search buffer and tokens are output, the size of the offset field should be increased. When the search buffer is between 25% and 50% full, the size of the offset should still be s - 1 and should be increased to s only when the buffer is 50% full or more. Even this sophisticated scheme can be improved upon to yield slightly better compression. Section 2.9 introduces phased-in codes and shows how to construct a set of pointers of two sizes to address an array of n elements even when n is not a power of 2. LZB employs phased-in codes to encode the size of the offset field at any point in the encoding. As an example, when there are nine symbols in the search buffer, LZB uses the seven 3-bit codes and two 4-bit codes 000, 001, 010, 011, 100, 101, 110, 1110, and 1111. The second field of an LZSS token is the match length l. This is normally a fairly small integer, but can sometimes have large values. Recall that LZSS computes a value p and outputs a token only if it requires fewer bits than emitting p raw characters. Thus, the smallest value of l is p+1. LZB employs the Elias gamma code of Section 3.4 to encode the length field. A match of length i is encoded as the gamma code of the integer i - p + 1. The length of the gamma code of the integer n is 1 + 2log2 n and it is ideal for cases where n appears in the input with probability 1/(2n2). As a simple example, assuming that p = 3, a match of five symbols is encoded as the gamma code of 5 - 3 + 1 = 3 which is 011. 6.4.2 SLH SLH is another variant of LZSS, described in [Brent 87]. It employs a circular buffer and it outputs raw symbols and pairs as does LZSS. The main difference is that SLH is a two-pass algorithm where the first pass employs a hash table to locate the best match and to count frequencies, and the second pass encodes the offsets and the raw symbols with Huffman codes prepared from the frequencies counted by the first pass. We start with a few details about the hash table H. The first pass starts with an empty H. As strings are searched, matches are either found in H or are not found and are then added to H. What is actually stored in a location of H when a string is added, is a pointer to the string in the search buffer. In order to find a match for a string s of symbols, s is passed through a hashing function that returns a pointer to H. Thus, a match to s can be located (or verified as not found) in H in one step. (This discussion of hash tables ignores collisions, and detailed descriptions of this data structure can be found in many texts.) We denote by p the break-even point, where a token representing a match of p or more symbols is shorter than the raw symbols themselves. At a general point in the encoding process, the search buffer has symbols a1 through aj and the look-ahead buffer has symbols aj+1 through an. We select the p-symbol string s = aj+1aj+2 . . . aj+p-1, hash it to a pointer q and check H[q]. If a match is not found, then s is added to H (i.e., an offset P is stored in H and is updated as the buffers are shifted and s moves to the 6.4 LZSS 343 left). If a match is found, the offset P found in H and the match length are prepared as the token pair (P, p) representing the best match so far. This token is saved but is not immediately output. In either case, match found or not found, the match length is incremented by 1 and the process is repeated in an attempt to find a longer match. In practice, the search buffer is implemented as a circular queue. Each time it is shifted, symbols are moved out of its “left” end and matches containing these symbols have to be deleted from H, which further complicates this algorithm. The second pass of SLH encodes the offsets and the raw characters with Huffman codes. The frequencies counted in the first pass are first normalized to the interval [0, 511] and a table with 512 Huffman codes is prepared (this table has to be written on the output, but its size is only a few hundred bytes). The implementation discussed in [Brent 87] assumes that the data symbols are bytes and that S, the size of the search buffer, is less than 218. Thus, an offset (a pointer to the search buffer) is at most 18 bits long. An 18-bit offset is broken up into two 9-bit parts and each part is Huffman encoded separately. Each raw character is also Huffman encoded. The developer of this method does not indicate how an 18-bit offset should be split in two and precisely what frequencies are counted by the first pass. If that pass counts the frequencies of the raw input characters, then these characters will be efficiently encoded by the Huffman codes, but the same Huffman codes will not correspond to the frequencies of the 9-bit parts of the offset. It is possible to count the frequencies of the 9-bit parts and prepare two separate Huffman code tables for them, but other LZ methods offer the same or better compression performance, while being simpler to implement and evaluate. The developer of this method does not explain the name SLH. The excellent book Text Compression [Bell et al. 90] refers to this method as LZH, but our book already uses LZH as the name of an LZ77 variant proposed by Bernd Herd, which leads to ambiguity. Even data compression experts cannot always avoid confusion. An expert is a man who tells you a simple thing in a confused way in such a fashion as to make you think the confusion is your own fault. —William Castle 6.4.3 LZARI The following is quoted from [Okumura 98]. During the summer of 1988, I [Haruhiko Okumura] wrote another compression program, LZARI. This program is based on the following observation: Each output of LZSS is either a single character or a position,length pair. A single character can be coded as an integer between 0 and 255. As for the length field, if the range of length is 2 to 257, say, it can be coded as an integer between 256 and 511. Thus, I can say that there are 512 kinds of “characters,” and the “characters” 256 through 511 are accompanied by a position field. These 512 “characters” can be Huffmancoded, or better still, algebraically coded. The position field can be coded in the same manner. In LZARI, I used an adaptive algebraic compression to encode the “characters,” and static algebraic compression to encode the position field. (There were several versions of LZARI; some of them were slightly different from the above description.) The compression of LZARI was very tight, though rather slow. 344 6. Dictionary Methods 6.5 LZPP The original LZ methods were published in 1977 and 1978. They established the field of dictionary-based compression, which is why the 1980s were the golden age of this approach to compression, Most LZ algorithms were developed during this decade, but LZPP, the topic of this section, is an exception. This method [Pylak 03] is a modern, sophisticated algorithm that extends LZSS in several directions and has been inspired by research done and experience gained by many workers in the 1990s. LZPP identifies several sources of redundancy in the various quantities generated and manipulated by LZSS and exploits these sources to obtain better overall compression. Virtually all LZ methods output offsets (also referred to as pointers or indexes) to a dictionary, and LZSS is no exception. However, many experiments indicate that the distribution of index values is not uniform and that most indexes are small, as illustrated by Figure 6.6. Thus, the set of indexes has large entropy (for the concatenated Calgary Corpus, [Pylak 03] gives this entropy as 8.954), which implies that the indexes can be compressed efficiently with an entropy encoder. LZPP employs a modified range encoder (Section 5.10.1) for this purpose. The original range coder has been modified by adding a circular buffer, an escape symbol, and an exclusion mechanism. In addition, each of the bytes of an index is compressed separately, which simplifies the range coder and allows for 3-byte indexes (i.e., a 16 MB dictionary).                             Figure 6.6: Distribution of Typical LZ Index Values (After [Pylak 03]). LZSS outputs either raw characters or pairs (match length, index). It turns out that the distribution of match lengths is also far from uniform, which makes it possible to compress them too with an entropy encoder. Figure 6.7 shows the distribution of match lengths for the small, 21 KB file obj1, part of the Calgary Corpus. It is clear that short matches, of three and four characters, are considerably more common than longer matches, thereby resulting in large entropy of match lengths. The same range coder is used by LZPP to encode the lengths, allowing for match lengths of up to 64 KB. The next source of redundancy in LZSS is the distribution of flag values. Recall that LZSS associates a flag with each raw character (a 0 flag) and each pair (a flag of 1) it outputs. Flag values of 2 and 3 are also used and are described below. The 6.5 LZPP 345                                             !"#$ %! !" Figure 6.7: Distribution of Match Lengths (After [Pylak 03]). redundancy in flag values exists because the probabilities of the two values are not 0.5 but vary considerably during encoding. It is intuitively clear that when the encoder starts, few matches exist, many raw characters are output, and most flags are therefore 0. As encoding progresses, more and more matches are found, so more flags of 1 are output. Thus, the probability of a 0 flag starts high, and slowly drops to a low value, hopefully much less than 0.5. This probability is illustrated in Figure 6.8, which shows how the probability of a 0 flag drops from almost 1 at the start, to around 0.2 after about 36,000 characters of the input have been read and processed. The input for the figure was file paper1 of the Calgary Corpus.                                  Figure 6.8: Probability of a 0 Flag (After [Pylak 03]). LZPP uses the same range encoder to compress the flags, which improves its compression pereformance by up to 2%. The description in this section shows that LZPP employs four different flags (as opposed to the two flags employed by LZSS), which makes the compression of flags even more meaningful. Another source of redundancy in LZSS is the raw characters. LZSS outputs raw characters, while LZPP compresses these characters with the same range encoder. 346 6. Dictionary Methods It should be mentioned that the range encoder employed by LZPP is a variant of adaptive arithmetic coding. Its main input is an item to be encoded, but it must also receive a probability for the item. LZPP uses order-0 and order-1 contexts to estimate the probabilities of the indexes, flags, and characters it sends to the range encoder. The concept of contexts and their various orders is covered in Section 5.14, but here is what order-0 context means. Suppose that the current raw character is q. LZPP counts the number of times each character has been encoded and emitted as a raw character token. If q has been encoded and output 35 times as a raw character token, then the order-0 probability estimated for this occurrence of q is 35/t, where t is the total number of raw character tokens output so far. Figure 6.7 shows that match lengths of 3 are much more common than any other lengths, which is why LZPP treats 3-byte matches differently. It constructs and maintains a special data structure (a FIFO queue) D that holds a certain number of 3-byte matches, with the number of occurrences of each. When a new 3-byte match is found in the dictionary, the encoder searches D. If the match is already there, the encoder outputs the single item index3, where index3 is the index to the match in D. A flag of 2 is associated with this item. An index to D is encoded by an entropy coder before it is output, and the important point is that the occurrence count of the match is used by LZPP to compute a probability which is sent to the entropy coder together with index3. The queue D is also exploited for a further small performance improvement. If LZPP outputs two consecutive raw characters, say, t and h, it knows that the next character x (the leftmost one in the look-ahead buffer) cannot be the last character of any of the 3-byte matches thy in D. Thus, if the trigrams the, tha, thi, and tho are in D, the next character cannot be e, a, or i. These characters are excluded when the x is encoded. When the encoder computes the probability of the x, it excludes the frequency counts for e, a, and i. The following illustrates another type of exclusion. If the search and look-ahead buffers hold the following characters: ‹ coded text. . . ... whiletheselet|themselvesout.... . .‹ text to be read and the strings the in the two buffers are matched, then LZPP knows that the character following the in the look-ahead buffer cannot be an s. Thus, the symbol s is excluded when LZPP computes a probability for encoding the length and index for string the. It has already been mentioned that LZPP handles 3-byte matches in a special way. Experiments with LZPP indicate that other short matches, those of four and five bytes, should also be treated differently. Specifically, it has been noticed that when a 4-byte or a 5-byte match is distant (i.e., has a large index), the resulting pair (length, index) does not yield good compression. As a result, LZPP limits the indexes for short matches. If the smallest index of a 4-byte match exceeds 255, the match is ignored. Similarly, 5-byte matches are ignored if the smallest index exceeds 65,535. The last source of redundancy in LZSS that is addressed by LZPP is symbol context. The concept of context is central to the PPM algorithm and is treated in detail in Section 5.14. The reader is advised to read the first few paragraphs of that section before continuing. The idea is that characters read from the input file are not random. Certain characters (such as e, t, and a in English) occur more often than others (such as j, q, and z). 6.5 LZPP 347 Also certain pairs and triplets (digrams and trigrams) of characters are common while others are rare. LZPP employs order-1 contexts to encode raw characters. It maintains a two-dimensional array C of 256Χ256 entries where it records the occurrences of digrams. Thus, if the leftmost character in the look-ahead buffer is h and there are no 3-byte or longer matches for it, LZPP is supposed to output the h as a raw character (suitably encoded). Before it does that, it locates the rightmost character in the search buffer, say, t, and examines the frequency count in row t column h of array C. If the count c found there (the number of previous occurrences of the digram th) is nonzero, the h is encoded with probability c/a, where a is the number of characters read so far from the input file. A flag of 3 is associated with the resulting bits. If c is zero (the digram th hasn’t been seen so far), the h is encoded and is output with a 0 flag. This is how LZPP handles the famous zero-probability problem in order-1 contexts. The escape mechanism. An escape character is added to the alphabet, and is used to encode those raw characters that have zero frequencies. LZPP maintains two counts for each character of the alphabet. The main count is maintained in array C. It is the number of occurrences of the character so far and it may be zero. The secondary count is 1 plus the number of times the character has been coded with the escape mechanism. The secondary count is always nonzero and is used when a character is encoded with the escape. Notice that this count is maintained only for the most-recent 4,096 characters encoded with an escape. When a character with zero main count is to be emitted, the escape flag is encoded first, followed by the character itself, encoded with a special probability that is computed by considering only the characters with zero occurrences and adding their secondary counts. A 64 KB hash table is used by the encoder to locate matches in the dictionary. The hash function is the well-known CRC-32 (Section 6.32). The leftmost four characters in the look-ahead buffer are considered a 32-bit integer and the value returned by CRC-32 for this integer is truncated to 16 bits and becomes an index to the 64 MB dictionary. Further search in the dictionary is needed to find matches longer than four bytes. Tests performed on the Calgary Corpus comparing several compression methods indicate that the compression ratios produced by LZPP are about 30% better than those of LZSS and about 1.5–2% worse than those of WinRAR (Section 6.22), a commercial product which is also based on LZSS. The developer of LZPP also notes that compression algorithms (including the commercial RK software) that are based on PPMZ (Section 5.14.7) outperform LZPP as well as any other dictionary-based algorithm. 6.5.1 Deficiencies Before we discuss LZ78, let’s summarize the deficiencies of LZ77 and its variants. It has already been mentioned that LZ77 uses the built-in implicit assumption that patterns in the input data occur close together. Data streams that don’t satisfy this assumption compress poorly. A common example is text where a certain word, say economy, occurs often but is uniformly distributed throughout the text. When this word is shifted into the look-ahead buffer, its previous occurrence may have already been shifted out of the search buffer. A better algorithm would save commonly-occurring strings in the dictionary and not simply slide it all the time. 348 6. Dictionary Methods Another disadvantage of LZ77 is the limited size L of the look-ahead buffer. The size of matched strings is limited to L-1, but L must be kept small because the process of matching strings involves comparing individual symbols. If L were doubled in size, compression would improve, since longer matches would be possible, but the encoder would be much slower when searching for long matches. The size S of the search buffer is also limited. A large search buffer results in better compression but slows down the encoder, because searching takes longer (even with a binary search tree). Increasing the sizes of the two buffers also means creating longer tokens, thereby reducing compression efficiency. With two-byte tokens, compressing a two-character string into one token results in two bytes plus a flag. Writing the two characters as two raw ASCII codes results in two bytes plus two flags, a very small difference in size. The encoder should, in such a case, use the latter choice and write the two characters in uncompressed form, saving time and wasting just one bit. We say that the encoder has a two-byte breakeven point. With longer tokens, the breakeven point increases to three bytes. 6.6 Repetition Times Frans Willems, one of the developers of context-tree weighting (Section 5.16), is also the developer of this original (although not very efficient) dictionary-based method. The input may consist of any symbols, but the method is described here and also in [Willems 89] for binary input. The input symbols are grouped into words of length L each that are placed in a sliding buffer. The buffer is divided into a look-ahead buffer with words still to be compressed, and a search buffer containing the B most-recently processed words. The encoder tries to match the leftmost word in the look-ahead buffer to the contents of the search buffer. Only one word in the look-ahead buffer is matched in each step. If a match is found, the distance (offset) of the word from the start of the match is denoted by m and is encoded by a 2-part prefix code that’s written on the compressed stream. Notice that there is no need to encode the number of symbols matched, because exactly one word is matched. If no match is found, a special code is written, followed by the L symbols of the unmatched word in raw format. The method is illustrated by a simple example. We assume that the input symbols are bits. We select L = 3 for the length of words, and a search buffer of length B = 2L - 1 = 7 containing the seven most-recently processed bits. The look-ahead buffer contains just the binary data, and the commas shown here are used only to indicate word boundaries. ‹ coded input. . . 0100100|100,000,011,111,011,101,001. . .‹ input to be read It is obvious that the leftmost word “100” in the look-ahead buffer matches the rightmost three bits in the search buffer. The repetition time (the offset) for this word is therefore m = 3. (The biggest repetition time is the length B of the search buffer, 7 in our example.) The buffer is now shifted one word (three bits) to the left to become ‹ . . . 0100100100|000,011,111,011,101,001,. . . . . .‹ input to be read The repetition time for the current word “000” is m = 1 because each bit in this word is matched with the bit immediately to its left. Notice that it is possible to match 6.6 Repetition Times 349 the leftmost 0 of the next word “011” with the bit to its left, but this method matches exactly one word in each step. The buffer is again shifted L positions to become ‹ . . . 0100100100000|011,111,011,101,001,. . . . . . . . .‹ input to be read There is no match for the next word “011” in the search buffer, so m is set to a special value that we denote by 8* (meaning; greater than or equal 8). It is easy to verify that the repetition times of the remaining three words are 6, 4, and 8*. Each repetition time is encoded by first determining two integers p and q. If m = 8*, then p is set to L; otherwise p is selected as the integer that satisfies 2p . m < 2p+1. Notice that p is located in the interval [0, L - 1]. The integer q is determined by q = m - 2p, which places it in the interval [0, 2p - 1]. Table 6.9 lists the values of m, p, q, and the prefix codes used for L = 3. m p q Prefix Suffix Length 1 0 0 00 none 2 2 1 0 01 0 3 3 1 1 01 1 3 4 2 0 10 00 4 5 2 1 10 01 4 6 2 2 10 10 4 7 2 3 10 11 4 8* 3 — 11 word 5 Table 6.9: Repetition Time Encoding Table for L = 3. Once p and q are known, a prefix code for m is constructed and is written on the compressed stream. It consists of two parts, a prefix and a suffix, that are the binary values of p and q, respectively. Since p is in the interval [0, L - 1], the prefix requires log(L + 1) bits. The length of the suffix is p bits. The case p = L is different. Here, the suffix is the raw value (L bits) of the word being compressed. The compressed stream for the seven words of our example consists of the seven codes 01|1, 00, 11|011, 00, 10|10, 10|00, 11|001, . . . , where the vertical bars separate the prefix and suffix of a code. Notice that the third and seventh words (011 and 001) are included in the codes in raw format. It is easy to see why this method generates prefix codes. Once a code has been assigned (such as 01|0, the code of m = 2), that code cannot be the prefix of any other code because (1) some of the other codes are for different values of p and thus do not start with 01, and (2) codes for the same p do start with 01 but must have different values of q, so they have different suffixes. The compression performance of this method is inferior to that of LZ77, but it is interesting for the following reasons. 1. It is universal and optimal. It does not use the statistics of the input stream, and its performance asymptotically approaches the entropy of the input as the input stream gets longer. 350 6. Dictionary Methods 2. It is shown in [Cachin 98] that this method can be modified to include data hiding (steganography). 6.7 QIC-122 QIC was an international trade association, incorporated in 1987, whose mission was to encourage and promote the widespread use of quarter-inch tape cartridge technology (hence the acronym QIC; see also http://www.qic.org/html). The association ceased to exist in 1998, when the technology it promoted became obsolete. The QIC-122 compression standard is an LZ77 variant that has been developed by QIC for text compression on 1/4-inch data cartridge tape drives. Data is read and shifted into a 2048-byte (= 211) input buffer from right to left, such that the first character is the leftmost one. When the buffer is full, or when all the data has been read into it, the algorithm searches from left to right for repeated strings. The output consists of raw characters and of tokens that represent strings already seen in the buffer. As an example, suppose that the following data have been read and shifted into the buffer: ABAAAAAACABABABA. . . . . . . . . . . . . . . The first character A is obviously not a repetition of any previous string, so it is encoded as a raw (ASCII) character (see below). The next character B is also encoded as raw. The third character A is identical to the first character but is also encoded as raw since repeated strings should be at least two characters long. Only with the fourth character A we do have a repeated string. The string of five A’s from position 4 to position 8 is identical to the one from position 3 to position 7. It is therefore encoded as a string of length 5 at offset 1. The offset in this method is the distance between the start of the repeated string and the start of the original one. The next character C at position 9 is encoded as raw. The string ABA at positions 10–12 is a repeat of the string at positions 1–3, so it is encoded as a string of length 3 at offset 10 - 1 = 9. Finally, the string BABA at positions 13–16 is encoded with length 4 at offset 2, since it is a repetition of the string at positions 10–13.  Exercise 6.5: Suppose that the next four characters of data are CAAC ABAAAAAACABABABACAAC. . . . . . . . . . . . How will they be encoded? A raw character is encoded as 0 followed by the 8 ASCII bits of the character. A string is encoded as a token that starts with 1 followed by the encoded offset, followed by the encoded length. Small offsets are encoded as 1, followed by 7 offset bits; large offsets are encoded as 0 followed by 11 offset bits (recall that the buffer size is 211). The length is encoded according to Table 6.10. The 9-bit string 110000000 is written, as an end marker, at the end of the output stream. 6.7 QIC-122 351 Bytes Length 2 00 3 01 4 10 5 11 00 6 11 01 7 11 10 8 11 11 0000 9 11 11 0001 10 11 11 0010 11 11 11 0011 12 11 11 0100 13 11 11 0101 14 11 11 0110 15 11 11 0111 16 11 11 1000 Bytes Length 17 11 11 1001 18 11 11 1010 19 11 11 1011 20 11 11 1100 21 11 11 1101 22 11 11 1110 23 11 11 1111 0000 24 11 11 1111 0001 25 11 11 1111 0010 ... 37 11 11 1111 1110 38 11 11 1111 1111 0000 39 11 11 1111 1111 0001 etc. Table 6.10: Values of the Field.  Exercise 6.6: How can the decoder identify the end marker? When the search algorithm arrives at the right end of the buffer, it shifts the buffer to the left and inputs the next character into the rightmost position of the buffer. The decoder is the reverse of the encoder (symmetric compression). When I saw each luminous creature in profile, from the point of view of its body, its egglike shape was like a gigantic asymmetrical yoyo that was standing edgewise, or like an almost round pot that was resting on its side with its lid on. The part that looked like a lid was the front plate; it was perhaps one-fifth the thickness of the total cocoon. —Carlos Castaneda, The Fire From Within (1984) Figure 6.11 is a precise description of the compression process, expressed in BNF, which is a metalanguage used to describe processes and formal languages unambiguously. BNF uses the following metasymbols: ::= The symbol on the left is defined by the expression on the right. An expression still to be defined. | A logical OR. [] Optional. The expression in the brackets may occur zero or more times. () A comment. 0,1 The bits 0 and 1. (Special applications of BNF may require more symbols.) Table 6.12 shows the results of encoding ABAAAAAACABABABA (a 16-symbol string). The reader can easily verify that the output stream consists of the 10 bytes 20 90 88 38 1C 21 E2 5C 15 80. 352 6. Dictionary Methods (QIC-122 BNF Description) ::=[] ::= 0 | 1 ::= (8-bit byte) ::= ::= 1 (a 7-bit offset) | 0 (an 11-bit offset) ::= (as per length table) ::=110000000 (Compressed bytes with offset=0) ::=0|1 Figure 6.11: BNF Definition of QIC-122. Raw byte “A” 0 01000001 Raw byte “B” 0 01000010 Raw byte “A” 0 01000001 String “AAAAA” offset=1 1 1 0000001 1100 Raw byte “C” 0 01000011 String “ABA” offset=9 1 1 0001001 01 String “BABA” offset=2 1 1 0000010 10 End-Marker 1 1 0000000 Table 6.12: Encoding the Example. 6.8 LZX In 1995, Jonathan Forbes and Tomi Poutanen developed an LZ variant (possibly influenced by Deflate) that they dubbed LZX [LZX 09]. The main feature of the first version of LZX was the way it encoded the match offsets, which can be large, by segmenting the size of the search buffer. They implemented LZX on the Amiga personal computer and included a feature where data was grouped into large blocks instead of being compressed as a single unit. At about the same time, Microsoft devised a new installation media format that it termed, in analogy to a file cabinet, cabinet files. A cabinet file has an extension name of .cab and may consist of several data files concatenated into one unit and compressed. Initially, Microsoft used two compression methods to compress cabinet files, MSZIP (which is just another name for Deflate) and Quantum, a large-window dictionary-based encoder that employs arithmetic coding. Quantum was developed by David Stafford. Microsoft later used cabinet files in its Cabinet Software Development Kit (SDK). This is a software package that provides software developers with the tools required to employ cabinet files in any applications that they implement. In 1997, Jonathan Forbes went to work for Microsoft and modified LZX to compress cabinet files. Microsoft published an official specification for cabinet files, including MSZIP and LZX, but excluding Quantum. The LZX description contained errors to 6.8 LZX 353 such an extent that it wasn’t possible to create a working implementation from it. LZX documentation is available in executable file Cabsdk.exe located at http:// download.microsoft.com/download/platformsdk/cab/2.0/w98nt42kmexp/en-us/. After unpacking this executable file, the documentation is found in file LZXFMT.DOC. LZX is also used to compress chm files. Microsoft Compiled HTML (chm) is a proprietary format initially developed by Microsoft (in 1997, as part of its Windows 98) for online help files. It was supposed to be the successor to the Microsoft WinHelp format. It seems that many Windows users liked this compact file format and started using it for all types of files, as an alternative to PDF. The popular lit file format is an extension of chm. A chm file is identified by a .chm file name extension. It consists of web pages written in a subset of HTML and a hyperlinked table of contents. A chm file can have a detailed index and table-of-contents, which is one reason for the popularity of this format. Currently, there are chm readers available for both Windows and Macintosh. LZX is a variant of LZ77 that writes on the compressed stream either unmatched characters or pairs (offset, length). What is actually written on the compressed stream is variable-length codes for the unmatched characters, offsets, and lengths. The size of the search buffer is a power of 2, between 215 and 221. LZX uses static canonical Huffman trees (Section 5.2.8) to provide variable-length, prefix codes for the three types of data. There are 256 possible character values, but there may be many different offsets and lengths. Thus, the Huffman trees have to be large, but any particular cabinet file being compressed by LZX may need just a small part of each tree. Those parts are written on the compressed stream. Because a single cabinet file may consist of several data files, each is compressed separately and is written on the compressed stream as a block, including those parts of the trees that it requires. The other important features of LZX are listed here. Repeated offsets. The developers of LZX noticed that certain offsets tend to repeat; i.e., if a certain string is compressed to a pair (74, length), then there is a good chance that offset 74 will be used again soon. Thus, the three special codes 0, 1, and 2 were allocated to encode three of the most-recent offsets. The actual offset associated with each of those codes varies all the time. We denote by R0, R1, and R2 the mostrecent, second most-recent, and third most-recent offsets, respectively (these offsets must themselves be nonrepeating; i.e., none should be 0, 1, or 2). We consider R0, R1, and R2 a short list and update it similar to an LRU (least-recently used) queue. The three quantities are initialized to 1 and are updated as follows. Assume that the current offset is X, then if X = R0 and X = R1 and X = R2, then R2 ‹ R1, R1 ‹ R0, R0 ‹ X, if X = R0, then nothing, if X = R1, then swap R0 and R1, if X = R2, then swap R0 and R2. Because codes 0, 1, and 2 are allocated to the three special recent offsets, an offset of 3 is allocated code 5, and, in general, an offset x is assigned code x+2. The largest offset is the size of the search buffer minus 3, and its assigned code is the size of the search buffer minus 1. 354 6. Dictionary Methods Encoder preprocessing. LZX was designed to compress Microsoft cabinet files, which are part of the Windows operating system. Computers using this system are generally based on microprocessors made by Intel, and throughout the 1980s and 1990s, before the introduction of the pentium, these microprocessors were members of the wellknown 80x86 family. The encoder preprocessing mode of LZX is selected by the user when the input stream is an executable file for an 80x86 computer. This mode converts 80x86 CALL instructions to use absolute instead of relative addresses. Output block format. LZX outputs the compressed data in blocks, where each block contains raw characters, offsets, match lengths, and the canonical Huffman trees used to encode the three types of data. A canonical Huffman tree can be reconstructed from the path length of each of its nodes. Thus, only the path lengths have to be written on the output for each Huffman tree. LZX limits the depth of a Huffman tree to 16, so each tree node is represented on the output by a number in the range 0 to 16. A 0 indicates a missing node (a Huffman code that’s not used by this block). If the tree has to be bigger, the current block is written on the output and compression resumes with fresh trees. The tree nodes are written in compressed form. If several consecutive tree nodes are identical, then run-length encoding is used to encode them. The three numbers 17, 18, and 19 are used for this purpose. Otherwise the difference (modulo 17) between the path lengths of the current node and the previous node is written. This difference is in the interval [0, 16]. Thus, what’s written on the output are the 20 5-bit integers 0 through 19, and these integers are themselves encoded by a Huffman tree called a pre-tree. The pre-tree is generated dynamically according to the frequencies of occurrence of the 20 values. The pre-tree itself has to be written on the output, and it is written as 20 4-bit integers (a total of 80 bits) where each integer indicates the path length of one of the 20 tree nodes. A path of length zero indicates that one of the 20 values is not used in the current block. The offsets and match lengths are themselves compressed in a complex process that involves several steps and is summarized in Figure 6.13. The individual steps involve many operations and use several tables that are built into both encoder and decoder. However, because LZX is not an important compression method, these steps are not discussed here. 6.9 LZ78 The LZ78 method (which is sometimes referred to as LZ2) [Ziv and Lempel 78] does not use any search buffer, look-ahead buffer, or sliding window. Instead, there is a dictionary of previously encountered strings. This dictionary starts empty (or almost empty), and its size is limited only by the amount of available memory. The encoder outputs two-field tokens. The first field is a pointer to the dictionary; the second is the code of a symbol. Tokens do not contain the length of a string, since this is implied in the dictionary. Each token corresponds to a string of input symbols, and that string is added to the dictionary after the token is written on the compressed stream. Nothing is ever deleted from the dictionary, which is both an advantage over LZ77 (since future strings can be compressed even by strings seen in the distant past) and a liability (because the dictionary tends to grow fast and to fill up the entire available memory). 6.9 LZ78 355 Match length Length footer Length tree Length/position Main tree Verbatim position bits bits Length header Position slot Formatted offset Match offset Position footer Aligned offset header Output tree Aligned offset Figure 6.13: LZX Processing of Offsets and Lengths. The dictionary starts with the null string at position zero. As symbols are input and encoded, strings are added to the dictionary at positions 1, 2, and so on. When the next symbol x is read from the input stream, the dictionary is searched for an entry with the one-symbol string x. If none are found, x is added to the next available position in the dictionary, and the token (0, x) is output. This token indicates the string “null x” (a concatenation of the null string and x). If an entry with x is found (at, say, position 37), the next symbol y is read, and the dictionary is searched for an entry containing the two-symbol string xy. If none are found, then string xy is added to the next available position in the dictionary, and the token (37, y) is output. This token indicates the string xy, since 37 is the dictionary position of string x. The process continues until the end of the input stream is reached. In general, the current symbol is read and becomes a one-symbol string. The encoder then tries to find it in the dictionary. If the symbol is found in the dictionary, the next symbol is read and concatenated with the first to form a two-symbol string that the encoder then tries to locate in the dictionary. As long as those strings are found in the dictionary, more symbols are read and concatenated to the string. At a certain point the string is not found in the dictionary, so the encoder adds it to the dictionary and outputs a token with the last dictionary match as its first field, and the last symbol of the string (the one that caused the search to fail) as its second field. Table 6.14 shows the first 14 steps in encoding the string sirsideastmaneasilyteasesseasickseals  Exercise 6.7: Complete Table 6.14. In each step, the string added to the dictionary is the one being encoded, minus 356 6. Dictionary Methods Dictionary Token Dictionary Token 0 null 1 “s” (0,“s”) 8 “a” (0,“a”) 2 “i” (0,“i”) 9 “st” (1,“t”) 3 “r” (0,“r”) 10 “m” (0,“m”) 4 “” (0,“”) 11 “an” (8,“n”) 5 “si” (1,“i”) 12 “ea” (7,“a”) 6 “d” (0,“d”) 13 “sil” (5,“l”) 7 “e” (4,“e”) 14 “y” (0,“y”) Table 6.14: First 14 Encoding Steps in LZ78. its last symbol. In a typical compression run, the dictionary starts with short strings, but as more text is being input and processed, longer and longer strings are added to it. The size of the dictionary can either be fixed or may be determined by the size of the available memory each time the LZ78 compression program is executed. A large dictionary may contain more strings and thus allow for longer matches, but the trade-off is longer pointers (and thus bigger tokens) and slower dictionary search. A good data structure for the dictionary is a tree, but not a binary tree. The tree starts with the null string as the root. All the strings that start with the null string (strings for which the token pointer is zero) are added to the tree as children of the root. In the above example those are s, i, r, , d, a, m, y, e, c, and k. Each of them becomes the root of a subtree as shown in Figure 6.15. For example, all the strings that start with s (the four strings si, sil, st, and s(eof)) constitute the subtree of node s. 8-a 25-l 11-n 17-s 22-c 16-e 20-a 18-s 6-d 2-i 23-k 10-m 3-r 1-s 14-y 26-eof 5-i 9-t 13-l 4- 19-s 7-e 15-t 24-e 21-i 12-a null Figure 6.15: An LZ78 Dictionary Tree. Assuming an alphabet with 8-bit symbols, there are 256 different symbols, so in principle, each node in the tree could have up to 256 children. The process of adding a child to a tree node should thus be dynamic. When the node is first created, it has no children and it should not reserve any memory space for them. As a child is added to the node, memory space should be claimed for it. Since no nodes are ever deleted, there is no need to reclaim memory space, which simplifies the memory management task somewhat. 6.9 LZ78 357 Such a tree makes it easy to search for a string and to add strings. To search for sil, for example, the program looks for the child s of the root, then for the child i of s, and so on, going down the tree. Here are some examples: 1. When the s of sid is input in step 5, the encoder finds node “1-s” in the tree as a child of “null”. It then inputs the next symbol i, but node s does not have a child i (in fact, it has no children at all at this point), so the encoder adds node “5-i” as a child of “1-s”, which effectively adds the string si to the tree. 2. When the blank space between eastman and easily is input in step 12, a similar situation happens. The encoder finds node “4-”, inputs e, finds “7-e”, inputs a, but “7-e” does not have “a” as a child, so the encoder adds node “12-a”, which effectively adds the string “ea” to the tree. A tree of the type described here is called a trie. In general, a trie is a tree in which the branching structure at any level is determined by just part of a data item, not the entire item (Section 5.14.5). In the case of LZ78, each string added to the tree effectively adds just one symbol, and does that by adding a branch. Since the total size of the tree is limited, it may fill up during compression. This, in fact, happens all the time except when the input stream is unusually small. The original LZ78 method does not specify what to do in such a case, so we list a few possible solutions. 1. The simplest solution is to freeze the dictionary at that point. No new nodes should be added, the tree becomes a static dictionary, but it can still be used to encode strings. 2. Delete the entire tree once it gets full and start with a new, empty tree. This solution effectively breaks the input into blocks, each with its own dictionary. If the content of the input varies from block to block, this solution will produce good compression, since it will eliminate a dictionary with strings that are unlikely to be used in the future. We can say that this solution implicitly assumes that future symbols will benefit more from new data than from old (the same implicit assumption used by LZ77). 3. The UNIX compress utility (Section 6.14) uses a more complex solution. 4. When the dictionary is full, delete some of the least-recently-used entries, to make room for new ones. Unfortunately there is no good algorithm to decide which entries to delete, and how many (but see the reuse procedure in Section 6.23). The LZ78 decoder works by building and maintaining the dictionary in the same way as the encoder. It is therefore more complex than the LZ77 decoder. Thus, LZ77 decoding is simpler than its encoding, but LZ78 encoding and decoding involve the same complexity. 358 6. Dictionary Methods 6.10 LZFG Edward Fiala and Daniel Greene have developed several related compression methods [Fiala and Greene 89] that are hybrids of LZ77 and LZ78. All their methods are based on the following scheme. The encoder generates a compressed file with tokens and literals (raw ASCII codes) intermixed. There are two types of tokens: a literal and a copy. A literal token indicates that a string of literals follows, a copy token points to a string previously seen in the data. The string theboyonmyrightistherightboy produces, when encoded, (literal 23)theboyonmyrightis(copy 4 23)(copy 6 13)(copy 3 29), where the three copy tokens refer to the strings the, right, and “boy”, respectively. The LZFG methods are best understood by considering how the decoder operates. The decoder starts with a large empty buffer in which it generates and shifts the decompressed stream. When the decoder inputs a (literal 23) token, it inputs the next 23 bytes as raw ASCII codes into the buffer, shifting the buffer such that the last byte input will be the rightmost one. When the decoder inputs (copy 4 23) it copies the string of length 4 that starts 23 positions from the right end of the buffer. The string is then appended to the buffer, while shifting it. Two LZFG variants, denoted by A1 and A2, are described here. The A1 scheme employs 8-bit literal tokens and 16-bit copy tokens. A literal token has the format 0000nnnn, where nnnn indicates the number of ASCII bytes following the token. Since the 4-bit nnnn field can have values between 0 and 15, they are interpreted as meaning 1 to 16. The longest possible string of literals is therefore 16 bytes. The format of a copy token is sssspp...p, where the 4-bit nonzero ssss field indicates the length of the string to be copied, and the 12-bit pp...p field is a displacement showing where the string starts in the buffer. Since the ssss field cannot be zero, it can have values only between 1 and 15, and they are interpreted as string lengths between 2 and 16. Displacement values are in the range [0, 4095] and are interpreted as [1, 4096]. The encoder starts with an empty search buffer, 4,096 bytes long, and fills up the look-ahead buffer with input data. At each subsequent step it tries to create a copy token. If nothing matches in that step, the encoder creates a literal token. Suppose that at a certain point the buffer contains ‹text already encoded.... ..xyz|abcd.. ....‹ text yet to be input The encoder tries to match the string abc... in the look-ahead buffer to various strings in the search buffer. If a match is found (of at least two symbols), a copy token is written on the compressed stream and the data in the buffers is shifted to the left by the size of the match. If a match is not found, the encoder starts a literal with the a and left-shifts the data one position. It then tries to match bcd... to the search buffer. If it finds a match, a literal token is output, followed by a byte with the a, followed by a match token. Otherwise, the b is appended to the literal and the encoder tries to match from cd... Literals can be up to 16 bytes long, so the string theboyonmy... above is encoded as (literal 16)theboyonmyri(literal 7)ghtis(copy 4 23)(copy 6 13)(copy 3 29). The A1 method borrows the idea of the sliding buffer from LZ77 but also behaves like LZ78, because it creates two-field tokens. This is why it can be considered a hybrid 6.10 LZFG 359 of the two original LZ methods. When A1 starts, it creates mostly literals, but when it gets up to speed (fills up its search buffer), it features strong adaptation, so more and more copy tokens appear in the compressed stream. The A2 method uses a larger search buffer (up to 21K bytes long). This improves compression, because longer copies can be found, but raises the problem of token size. A large search buffer implies large displacements in copy tokens; long copies imply large “length” fields in those tokens. At the same time we expect both the displacement and the “length” fields of a typical copy token to be small, since most matches are found close to the beginning of the search buffer. The solution is to use a variable-length code for those fields, and A2 uses the general unary codes of Section 3.1. The “length” field of a copy token is encoded with a (2,1,10) code (Table 3.4), making it possible to match strings up to 2,044 symbols long. Notice that the (2,1,10) code is between 3 and 18 bits long. The first four codes of the (2, 1, 10) code are 000, 001, 010, and 011. The last three of these codes indicate match lengths of two, three, and four, respectively (recall that the minimum match length is 2). The first one (code 000) is reserved to indicate a literal. The length of the literal then follows and is encoded with code (0, 1, 5). A literal can therefore be up to 63 bytes long, and the literal-length field in the token is encoded by between 1 and 10 bits. In case of a match, the “length” field is not 000 and is followed by the displacement field, which is encoded with the (10,2,14) code (Table 6.16). This code has 21K values, and the maximum code size is 16 bits (but see points 2 and 3 below). a = nth Number of Range of n 10+n · 2 codeword codewords integers 0 10 0x...x  10 210 = 1K 0–1023 1 12 10xx...x    12 212 = 4K 1024–5119 2 14 11xx...xx    14 214 = 16K 5120–21503 Total 21504 Table 6.16: The General Unary Code (10, 2, 14). Three more refinements are employed by the A2 method, to achieve slightly better (1% or 2%) compression. 1. A literal of maximum length (63 bytes) can immediately be followed by another literal or by a copy token of any length, but a literal of fewer than 63 bytes must be followed by a copy token matching at least three symbols (or by the end-of-file). This fact is used to shift down the (2,1,10) codes used to indicate the match length. Normally, codes 000, 001, 010, and 011 indicate no match, and matches of length 2, 3, and 4, respectively. However, a copy token following a literal token of fewer than 63 bytes uses codes 000, 001, 010, and 011 to indicate matches of length 3, 4, 5, and 6, respectively. This way the maximum match length can be 2,046 symbols instead of 2,044. 2. The displacement field is encoded with the (10, 2, 14) code, which has 21K values and whose individual codes range in size from 11 to 16 bits. For smaller files, such large 360 6. Dictionary Methods displacements may not be necessary, and other general unary codes may be used, with shorter individual codes. Method A2 thus uses codes of the form (10 - d, 2, 14 - d) for d = 10, 9, 8, . . . , 0. For d = 1, code (9, 2, 13) has 29 + 211 + 213 = 10,752 values, and individual codes range in size from 9 to 15 bits. For d = 10 code (0, 2, 4) contains 20 + 22 + 24 = 21 values, and codes are between 1 and 6 bits long. Method A2 starts with d = 10 [meaning it initially uses code (0, 2, 4)] and a search buffer of size 21 bytes. When the buffer fills up (indicating an input stream longer than 21 bytes), the A2 algorithm switches to d = 9 [code (1, 2, 5)] and increases the search buffer size to 42 bytes. This process continues until the entire input stream has been encoded or until d = 0 is reached [at which point code (10,2,14) is used to the end]. A lot of work for a small gain in compression! (See the discussion of diminishing returns (a word to the wise) in the Preface.) 3. Each of the codes (10 - d, 2, 14 - d) requires a search buffer of a certain size, from 21 up to 21K = 21,504 bytes, according to the number of codes it contains. If the user wants, for some reason, to assign the search buffer a different size, then some of the longer codes may never be used, which makes it possible to cut down a little the size of the individual codes. For example, if the user decides to use a search buffer of size 16K = 16,384 bytes, then code (10, 2, 14) has to be used [because the next code (9, 2, 13) contains just 10,752 values]. Code (10, 2, 14) contains 21K = 21,504 individual codes, so the 5,120 longest codes will never be used. The last group of codes (“11” followed by 14 bits) in (10, 2, 14) contains 214 = 16, 384 different individual codes, of which only 11,264 will be used. Of the 11,264 codes the first 8,192 can be represented as “11” followed by log2 11, 264 = 13 bits, and only the remaining 3,072 codes require log2 11, 264 = 14 bits to follow the first “11”. We thus end up with 8,192 15-bit codes and 3,072 16-bit codes, instead of 11,264 16-bit codes, a very small improvement. These three improvements illustrate the great lengths that researchers are willing to go to in order to improve their algorithms ever so slightly. Experience shows that fine-tuning an algorithm to squeeze out the last remaining bits of redundancy from the data gives diminishing returns. Modifying an algorithm to improve compression by 1% may increase the run time by 10% (from the Introduction). The LZFG “corpus” of algorithms contains four more methods. B1 and B2 are similar to A1 and A2 but faster because of the way they compute displacements. However, some compression ratio is sacrificed. C1 and C2 go in the opposite direction. They achieve slightly better compression than A1 and A2 at the price of slower operation. (LZFG has been patented, an issue that’s discussed in Section 6.34.) 6.11 LZRW1 361 6.11 LZRW1 Developed by Ross Williams [Williams 91a] and [Williams 91b] as a simple, fast LZ77 variant, LZRW1 is also related to method A1 of LZFG (Section 6.10). The main idea is to find a match in one step, using a hash table. This is fast but not very efficient, since the match found is not always the longest. We start with a description of the algorithm, follow with the format of the compressed stream, and conclude with an example. The method uses the entire available memory as a buffer and encodes the input stream in blocks. A block is read into the buffer and is completely encoded, then the next block is read and encoded, and so on. The length of the search buffer is 4K and that of the look-ahead buffer is 16 bytes. These two buffers slide along the input block in memory from left to right. Only one pointer, p_src, needs be maintained, pointing to the start of the look-ahead buffer. The pointer p_src is initialized to 1 and is incremented after each phrase is encoded, thereby moving both buffers to the right by the length of the phrase. Figure 6.17 shows how the search buffer starts empty, grows to 4K, and then starts sliding to the right, following the look-ahead buffer. search buffer (4096) already encoded look-ahead not encoded yet buffer (16) p-src look-ahead not encoded yet buffer (16) small search buffer p-src look-ahead not encoded yet buffer (16) initial p-src Figure 6.17: Sliding the LZRW1 Search- and Look-Ahead Buffers. The leftmost three characters of the look-ahead buffer are hashed into a 12-bit number I, which is used to index an array of 212 = 4,096 pointers. A pointer P is retrieved and is immediately replaced in the array by p_src. If P points outside the search buffer, there is no match; the first character in the look-ahead buffer is output as a literal, and p_src is advanced by 1. The same thing is done if P points inside the 362 6. Dictionary Methods search buffer but to a string that does not match the one in the look-ahead buffer. If P points to a match of at least three characters, the encoder finds the longest match (at most 16 characters), outputs a match item, and advances p_src by the length of the match. This process is depicted in Figure 6.19. An interesting point to note is that the array of pointers does not have to be initialized when the encoder starts, since the encoder checks every pointer. Initially, all pointers are random, but as they are replaced, more and more of them point to real matches. The output of the LZRW1 encoder (Figure 6.20) consists of groups, each starting with a 16-bit control word, followed by 16 items. Each item is either an 8-bit literal or a 16-bit copy item (a match) consisting of a 4-bit length field b (where the length is b+1) and a 12-bit offset (the a and c fields). The length field indicates lengths between 3 and 16. The 16 bits of the control word flag each of the 16 items that follow (a 0 flag indicates a literal and a flag of 1 indicates a match item). Obviously, groups have different lengths. The last group may contain fewer than 16 items. The decoder is even simpler than the encoder, since it does not need the array of pointers. It maintains a large buffer using a p_src pointer in the same way as the encoder. The decoder reads a control word from the compressed stream and uses its 16 bits to read 16 items. A literal item is decoded by appending it to the buffer and incrementing p_src by 1. A copy item is decoded by subtracting the offset from p_src, fetching a string from the search buffer, of length indicated by the length field, and appending it to the buffer. Then p_src is incremented by the length. Table 6.18 illustrates the first seven steps of encoding “thatthatchthaws”. The values produced by the hash function are arbitrary. Initially, all pointers are random (indicated by “any”) but they are replaced by useful ones very quickly.  Exercise 6.8: Summarize the last steps in a table similar to Table 6.18 and write the final compressed stream in binary. Hash p_src 3 chars index P Output Binary output 1 tha 4 any› 1 t 01110100 2 hat 6 any› 2 h 01101000 3 at 2 any›3 a 01100001 4 tt 1 any› 4 t 01110100 5 th 5 any› 5  00100000 6 tha 4 4›1 6,5 0000|0011|00000101 10 ch 3 any› 10 c 01100011 Table 6.18: First Seven Steps of Encoding that thatch thaws. Tests done by the original developer indicate that LZRW1 performs about 10% worse than LZC (the UNIX compress utility) but is four times faster. Also, it performs about 4% worse than LZFG (the A1 method) but runs ten times faster. It is therefore suited for cases where speed is more important than compression performance. A 68000 6.11 LZRW1 363 search buffer (4096) look-ahead buffer (16) p-src hash function offset initial p-src large buffer for one input block random pointer 4096 pointers already encoded not encoded yet Figure 6.19: The LZRW1 Encoder. 16-bit control word 16 items ASCII code a b c Compressed stream One group Figure 6.20: Format of the Output. 364 6. Dictionary Methods assembly language implementation has required, on average, the execution of only 13 machine instructions to compress, and four instructions to decompress, one byte.  Exercise 6.9: Show a practical situation where compression speed is more important than compression ratio. 6.12 LZRW4 LZRW4 is a variant of LZ77, based on ideas of Ross Williams about possible ways to combine a dictionary method with prediction (Section 6.35). LZRW4 also borrows some ideas from LZRW1. It uses a 1 Mbyte buffer where both the search and look-ahead buffers slide from left to right. At any point in the encoding process, the order-2 context of the current symbol (the two most-recent symbols in the search buffer) is used to predict the current symbol. The two symbols constituting the context are hashed to a 12-bit number I, which is used as an index to a 212 = 4,096-entry array A of partitions. Each partition contains 32 pointers to the input data in the 1 Mbyte buffer (each pointer is therefore 20 bits long). The 32 pointers in partition A[I] are checked to find the longest match between the look-ahead buffer and the input data seen so far. The longest match is selected and is coded in 8 bits. The first 3 bits code the match length according to Table 6.21; the remaining 5 bits identify the pointer in the partition. Such an 8-bit number is called a copy item. If no match is found, a literal is encoded in 8 bits. For each item, an extra bit is prepared, a 0 for a literal and a 1 for a copy item. The extra bits are accumulated in groups of 16, and each group is output, as in LZRW1, preceding the 16 items it refers to. 3 bits: 000 001 010 011 100 101 110 111 length: 2 3 4 5 6 7 8 16 Table 6.21: Encoding the Length in LZRW4. The partitions are updated all the time by moving “good” pointers toward the start of their partition. When a match is found, the encoder swaps the selected pointer with the pointer halfway toward the partition (Figure 6.22a,b). If no match is found, the entire 32-pointer partition is shifted to the left and the new pointer is entered on the right, pointing to the current symbol (Figure 6.22c). The Red Queen shook her head, “You may call it ‘nonsense’ if you like,” she said, “but I’ve heard nonsense, compared with which that would be as sensible as a dictionary!” —Lewis Carroll, Through the Looking Glass (1872) 6.13 LZW 365 l/2 partition (32 pointers) l swap search buffer look-ahead shift pointers new pointer (a) (b) (c) partition (32 pointers) Figure 6.22: Updating an LZRW4 Partition. 6.13 LZW This is a popular variant of LZ78, developed by Terry Welch in 1984 ([Welch 84] and [Phillips 92]). Its main feature is eliminating the second field of a token. An LZW token consists of just a pointer to the dictionary. To best understand LZW, we will temporarily forget that the dictionary is a tree, and will think of it as an array of variable-size strings. The LZW method starts by initializing the dictionary to all the symbols in the alphabet. In the common case of 8-bit symbols, the first 256 entries of the dictionary (entries 0 through 255) are occupied before any data is input. Because the dictionary is initialized, the next input character will always be found in the dictionary. This is why an LZW token can consist of just a pointer and does not have to contain a character code as in LZ77 and LZ78. (LZW has been patented and for many years its use required a license. This issue is treated in Section 6.34.) The principle of LZW is that the encoder inputs symbols one by one and accumulates them in a string I. After each symbol is input and is concatenated to I, the dictionary is searched for string I. As long as I is found in the dictionary, the process continues. At a certain point, adding the next symbol x causes the search to fail; string I is in the dictionary but string Ix (symbol x concatenated to I) is not. At this point the encoder (1) outputs the dictionary pointer that points to string I, (2) saves string Ix (which is now called a phrase) in the next available dictionary entry, and (3) initializes string I to symbol x. To illustrate this process, we again use the text string sirsideastmaneasilyteasesseasickseals. The steps are as follows: 0. Initialize entries 0–255 of the dictionary to all 256 8-bit bytes. 1. The first symbol s is input and is found in the dictionary (in entry 115, since this is 366 6. Dictionary Methods the ASCII code of s). The next symbol i is input, but si is not found in the dictionary. The encoder performs the following: (1) outputs 115, (2) saves string si in the next available dictionary entry (entry 256), and (3) initializes I to the symbol i. 2. The r of sir is input, but string ir is not in the dictionary. The encoder (1) outputs 105 (the ASCII code of i), (2) saves string ir in the next available dictionary entry (entry 257), and (3) initializes I to the symbol r. Table 6.23 summarizes all the steps of this process. Table 6.24 shows some of the original 256 entries in the LZW dictionary plus the entries added during encoding of the string above. The complete output stream is (only the numbers are output, not the strings in parentheses) as follows: 115 (s), 105 (i), 114 (r), 32 (), 256 (si), 100 (d), 32 (), 101 (e), 97 (a), 115 (s), 116 (t), 109 (m), 97 (a), 110 (n), 262 (e), 264 (as), 105 (i), 108 (l), 121 (y), 32 (), 116 (t), 263 (ea), 115 (s), 101 (e), 115 (s), 259 (s), 263 (ea), 259 (s), 105 (i), 99 (c), 107 (k), 280 (se), 97 (a), 108 (l), 115 (s), eof. Figure 6.25 is a pseudo-code listing of the algorithm. We denote by . the empty string, and by <> the concatenation of strings a and b. The line “append <> to the dictionary” is of special interest. It is clear that in practice, the dictionary may fill up. This line should therefore include a test for a full dictionary, and certain actions for the case where it is full. Since the first 256 entries of the dictionary are occupied right from the start, pointers to the dictionary have to be longer than 8 bits. A simple implementation would typically use 16-bit pointers, which allow for a 64K-entry dictionary (where 64K = 216 = 65,536). Such a dictionary will, of course, fill up very quickly in all but the smallest compression jobs. The same problem exists with LZ78, and any solutions used with LZ78 can also be used with LZW. Another interesting fact about LZW is that strings in the dictionary become only one character longer at a time. It therefore takes a long time to end up with long strings in the dictionary, and thus a chance to achieve really good compression. We can say that LZW adapts slowly to its input data.  Exercise 6.10: Use LZW to encode the string alfeatsalfalfa. Show the encoder output and the new entries added by it to the dictionary.  Exercise 6.11: Analyze the LZW compression of the string “aaaa...”. A dirty icon (anagram of “dictionary”) 6.13.1 LZW Decoding To understand how the LZW decoder works, we recall the three steps the encoder performs each time it writes something on the output stream. They are (1) it outputs the dictionary pointer that points to string I, (2) it saves string Ix in the next available entry of the dictionary, and (3) it initializes string I to symbol x. The decoder starts with the first entries of its dictionary initialized to all the symbols of the alphabet (normally 256 symbols). It then reads its input stream (which consists of pointers to the dictionary) and uses each pointer to retrieve uncompressed symbols from its dictionary and write them on its output stream. It also builds its dictionary in 6.13 LZW 367 in new in new I dict? entry output I dict? entry output s Y y Y si N 256-si 115 (s) y N 274-y 121 (y) i Y  Y ir N 257-ir 105 (i) t N 275-t 32 () r Y t Y r N 258-r 114 (r) te N 276-te 116 (t)  Y e Y s N 259-s 32 () ea Y s Y eas N 277-eas 263 (ea) si Y s Y sid N 260-sid 256 (si) se N 278-se 115 (s) d Y e Y d N 261-d 100 (d) es N 279-es 101 (e)  Y s Y e N 262-e 32 () s N 280-s 115 (s) e Y  Y ea N 263-ea 101 (e) s Y a Y se N 281-se 259 (s) as N 264-as 97 (a) e Y s Y ea Y st N 265-st 115 (s) ea N 282-ea 263 (ea) t Y  Y tm N 266-tm 116 (t) s Y m Y si N 283-si 259 (s) ma N 267-ma 109 (m) i Y a Y ic N 284-ic 105 (i) an N 268-an 97 (a) c Y n Y ck N 285-ck 99 (c) n N 269-n 110 (n) k Y  Y k N 286-k 107 (k) e Y  Y ea N 270-ea 262 (e) s Y a Y se Y as Y sea N 287-sea 281 (se) asi N 271-asi 264 (as) a Y i Y al N 288-al 97 (a) il N 272-il 105 (i) l Y l Y ls N 289-ls 108 (l) ly N 273-ly 108 (l) s Y s,eof N 115 (s) Table 6.23: Encoding sir sid eastman easily teases sea sick seals. 368 6. Dictionary Methods 0 NULL 110 n 262 e 276 te 1 SOH . . . 263 ea 277 eas . . . 115 s 264 as 278 se 32 SP 116 t 265 st 279 es . . . . . . 266 tm 280 s 97 a 121 y 267 ma 281 se 98 b . . . 268 an 282 ea 99 c 255 255 269 n 283 si 100 d 256 si 270 ea 284 ic 101 e 257 ir 271 asi 285 ck . . . 258 r 272 il 286 k 107 k 259 s 273 ly 287 sea 108 l 260 sid 274 y 288 al 109 m 261 d 275 t 289 ls Table 6.24: An LZW Dictionary. for i:=0 to 255 do append i as a 1-symbol string to the dictionary; append . to the dictionary; di:=dictionary index of .; repeat read(ch); if <> is in the dictionary then di:=dictionary index of <>; else output(di); append <> to the dictionary; di:=dictionary index of ch; endif; until end-of-input; Figure 6.25: The LZW Algorithm. 6.13 LZW 369 the same way as the encoder (this fact is usually expressed by saying that the encoder and decoder are synchronized, or that they work in lockstep). In the first decoding step, the decoder inputs the first pointer and uses it to retrieve a dictionary item I. This is a string of symbols, and it is written on the decoder’s output. String Ix needs to be saved in the dictionary, but symbol x is still unknown; it will be the first symbol in the next string retrieved from the dictionary. In each decoding step after the first, the decoder inputs the next pointer, retrieves the next string J from the dictionary, writes it on the output, isolates its first symbol x, and saves string Ix in the next available dictionary entry (after checking to make sure string Ix is not already in the dictionary). The decoder then moves J to I and is ready for the next step. In our sirsid... example, the first pointer that’s input by the decoder is 115. This corresponds to the string s, which is retrieved from the dictionary, gets stored in I, and becomes the first item written on the decoder’s output. The next pointer is 105, so string i is retrieved into J and is also written on the output. J’s first symbol is concatenated with I, to form string si, which does not exist in the dictionary, and is therefore added to it as entry 256. Variable J is moved to I, so I is now the string i. The next pointer is 114, so string r is retrieved from the dictionary into J and is also written on the output. J’s first symbol is concatenated with I, to form string ir, which does not exist in the dictionary, and is added to it as entry 257. Variable J is moved to I, so I is now the string r. The next step reads pointer 32, writes  on the output, and saves string r .  Exercise 6.12: Decode the string alfeatsalfalfa by using the encoding results from Exercise 6.10.  Exercise 6.13: Assume a two-symbol alphabet with the symbols a and b. Show the first few steps for encoding and decoding the string “ababab...”. 6.13.2 LZW Dictionary Structure Up until now, we have assumed that the LZW dictionary is an array of variable-size strings. To understand why a trie is a better data structure for the dictionary we need to recall how the encoder works. It inputs symbols and concatenates them into a variable I as long as the string in I is found in the dictionary. At a certain point the encoder inputs the first symbol x, which causes the search to fail (string Ix is not in the dictionary). It then adds Ix to the dictionary. This means that each string added to the dictionary effectively adds just one new symbol, x. (Phrased another way; for each dictionary string of more than one symbol, there exists a “parent” string in the dictionary that’s one symbol shorter.) A tree similar to the one used by LZ78 is therefore a good data structure, because adding string Ix to such a tree is done by adding one node with x. The main problem is that each node in the LZW tree may have many children (this is a multiway tree, not a binary tree). Imagine the node for the letter a in entry 97. Initially it has no children, but if the string ab is added to the tree, node 97 gets one child. Later, when, say, the string ae is added, node 97 gets a second child, and so on. The data structure for the tree should therefore be designed such that a node could have any number of children, but without having to reserve any memory for them in advance. 370 6. Dictionary Methods One way of designing such a data structure is to house the tree in an array of nodes, each a structure with two fields: a symbol and a pointer to the parent node. A node has no pointers to any child nodes. Moving down the tree, from a node to one of its children, is done by a hashing process in which the pointer to the node and the symbol of the child are hashed to create a new pointer. Suppose that string abc has already been input, symbol by symbol, and has been stored in the tree in the three nodes at locations 97, 266, and 284. Following that, the encoder has just input the next symbol d. The encoder now searches for string abcd, or, more specifically, for a node containing the symbol d whose parent is at location 284. The encoder hashes the 284 (the pointer to string abc) and the 100 (ASCII code of d) to create a pointer to some node, say, 299. The encoder then examines node 299. There are three possibilities: 1. The node is unused. This means that abcd is not yet in the dictionary and should be added to it. The encoder adds it to the tree by storing the parent pointer 284 and ASCII code 100 in the node. The result is the following: Node Address : 97 266 284 299 Contents : (-:a) (97:b) (266:c) (284:d) Represents: a ab abc abcd 2. The node contains a parent pointer of 284 and the ASCII code of d. This means that string abcd is already in the tree. The encoder inputs the next symbol, say e, and searches the dictionary tree for string abcde. 3. The node contains something else. This means that another hashing of a pointer and an ASCII code has resulted in 299, and node 299 already contains information from another string. This is called a collision, and it can be dealt with in several ways. The simplest way to deal with a collision is to increment pointer 299 and examine nodes 300, 301,. . . until an unused node is found, or until a node with (284:d) is found. In practice, we build nodes that are structures with three fields, a pointer to the parent node, the pointer (or index) created by the hashing process, and the code (normally ASCII) of the symbol contained in the node. The second field is necessary because of collisions. A node can therefore be illustrated by parent index symbol We illustrate this data structure using string ababab... of Exercise 6.13. The dictionary is an array dict where each entry is a structure with the three fields parent, index, and symbol. We refer to a field by, for example, dict[pointer].parent, where pointer is an index to the array. The dictionary is initialized to the two entries a and b. (To keep the example simple we use no ASCII codes. We assume that a has code 1 and b has code 2.) The first few steps of the encoder are as follows: Step 0: Mark all dictionary locations from 3 on as unused. / 1a / 2b / - / - / - . . . 6.13 LZW 371 Step 1: The first symbol a is input into variable I. What is actually input is the code of a, which in our example is 1, so I = 1. Since this is the first symbol, the encoder assumes that it is in the dictionary and so does not perform any search. Step 2: The second symbol b is input into J, so J = 2. The encoder has to search for string ab in the dictionary. It executes pointer:=hash(I,J). Let’s assume that the result is 5. Field dict[pointer].index contains “unused”, since location 5 is still empty, so string ab is not in the dictionary. It is added by executing dict[pointer].parent:=I; dict[pointer].index:=pointer; dict[pointer].symbol:=J; with pointer=5. J is moved into I, so I = 2. / 1a / 2b / - / - 1 5b . . . Step 3: The third symbol a is input into J, so J = 1. The encoder has to search for string ba in the dictionary. It executes pointer:=hash(I,J). Let’s assume that the result is 8. Field dict[pointer].index contains “unused”, so string ba is not in the dictionary. It is added as before by executing dict[pointer].parent:=I; dict[pointer].index:=pointer; dict[pointer].symbol:=J; with pointer=8. J is moved into I, so I = 1. / 1a / 2b / - / - 1 5b / - / - 2 8a / - . . . Step 4: The fourth symbol b is input into J, so J=2. The encoder has to search for string ab in the dictionary. It executes pointer:=hash(I,J). We know from step 2 that the result is 5. Field dict[pointer].index contains 5, so string ab is in the dictionary. The value of pointer is moved into I, so I = 5. Step 5: The fifth symbol a is input into J, so J = 1. The encoder has to search for string aba in the dictionary. It executes as usual pointer:=hash(I,J). Let’s assume that the result is 8 (a collision). Field dict[pointer].index contains 8, which looks good, but field dict[pointer].parent contains 2 instead of the expected 5, so the hash function knows that this is a collision and string aba is not in dictionary entry 8. It increments pointer as many times as necessary until it finds a dictionary entry with index=8 and parent=5 or until it finds an unused entry. In the former case, string aba is in the dictionary, and pointer is moved to I. In the latter case aba is not in the dictionary, and the encoder saves it in the entry pointed at by pointer, and moves J to I. / 1a / 2b / - / - 1 5b / - / - 2 8a 5 8a / - . . . Example: The 15 hashing steps for encoding the string alfeatsalfalfa are 372 6. Dictionary Methods shown below. The encoding process itself is illustrated in detail in the answer to Exercise 6.10. The results of the hashing are arbitrary; they are not the results produced by a real hash function. The 12 trie nodes constructed for this string are shown in Figure 6.26. 1. Hash(l,97) › 278. Array location 278 is set to (97, 278, l). 2. Hash(f,108) › 266. Array location 266 is set to (108, 266, f). 3. Hash(,102) › 269. Array location 269 is set to (102,269,). 4. Hash(e,32) › 267. Array location 267 is set to (32, 267, e). 5. Hash(a,101) › 265. Array location 265 is set to (101, 265, a). 6. Hash(t,97) › 272. Array location 272 is set to (97, 272, t). 7. Hash(s,116) › 265. A collision! Skip to the next available location, 268, and set it to (116, 265, s). This is why the index needs to be stored. 8. Hash(,115) › 270. Array location 270 is set to (115, 270, ). 9. Hash(a,32) › 268. A collision! Skip to the next available location, 271, and set it to (32, 268, a). 10. Hash(l,97) › 278. Array location 278 already contains index 278 and symbol l from step 1, so there is no need to store anything else or to add a new trie entry. 11. Hash(f,278) › 276. Array location 276 is set to (278, 276, f). 12. Hash(a,102) › 274. Array location 274 is set to (102, 274, a). 13. Hash(l,97) › 278. Array location 278 already contains index 278 and symbol l from step 1, so there is no need to do anything. 14. Hash(f,278) › 276. Array location 276 already contains index 276 and symbol f from step 11, so there is no need to do anything. 15. Hash(a,276) › 274. A collision! Skip to the next available location, 275, and set it to (276, 274, a). Readers who have carefully followed the discussion up to this point will be happy to learn that the LZW decoder’s use of the dictionary tree-array is simple and no hashing is needed. The decoder starts, like the encoder, by initializing the first 256 array locations. It then reads pointers from its input stream and uses each to locate a symbol in the dictionary. In the first decoding step, the decoder inputs the first pointer and uses it to retrieve a dictionary item I. This is a symbol that is now written by the decoder on its output stream. String Ix needs to be saved in the dictionary, but symbol x is still unknown; it will be the first symbol in the next string retrieved from the dictionary. In each decoding step after the first, the decoder inputs the next pointer and uses it to retrieve the next string J from the dictionary and write it on the output stream. If the pointer is, say 8, the decoder examines field dict[8].index. If this field equals 8, then this is the right node. Otherwise, the decoder examines consecutive array locations until it finds the right one. Once the right tree node is found, the parent field is used to go back up the tree and retrieve the individual symbols of the string in reverse order. The symbols are then placed in J in the right order (see below), the decoder isolates the first symbol x of J, and saves string Ix in the next available array location. (String I was found in the previous step, so only one node, with symbol x, needs be added.) The decoder then moves J to I and is ready for the next step. 6.13 LZW 373 265 266 267 268 269 270 271 272 273 274 275 276 277 278 / - / - / - / - / - / - / - / - / - / - / - / - / - 97 278 l / - 108 266 f / - / - / - / - / - / - / - / - / - / - / - 97 278 l / - 108 266 f / - / - 102 269  / - / - / - / - / - / - / - / - 97 278 l / - 108 266 f 32 267 e / - 102 269  / - / - / - / - / - / - / - / - 97 278 l 101 265 a 108 266 f 32 267 e / - 102 269  / - / - / - / - / - / - / - / - 97 278 l 101 265 a 108 266 f 32 267 e / - 102 269  / - / - 97 272 t / - / - / - / - / - 97 278 l 101 265 a 108 266 f 32 267 e 116 265 s 102 269  / - / - 97 272 t / - / - / - / - / - 97 278 l 101 265 a 108 266 f 32 267 e 116 265 s 102 269  115 270  / - 97 272 t / - / - / - / - / - 97 278 l 101 265 a 108 266 f 32 267 e 116 265 s 102 269  115 270  32 268 a 97 272 t / - / - / - / - / - 97 278 l 101 265 a 108 266 f 32 267 e 116 265 s 102 269  115 270  32 268 a 97 272 t / - / - / - 278 276 f / - 97 278 l 101 265 a 108 266 f 32 267 e 116 265 s 102 269  115 270  32 268 a 97 272 t / - 102 274 a / - 278 276 f / - 97 278 l 101 265 a 108 266 f 32 267 e 116 265 s 102 269  115 270  32 268 a 97 272 t / - 102 274 a 276 274 a 278 276 f / - 97 278 l Figure 6.26: Growing An LZW Trie for “alf eats alfalfa”. 374 6. Dictionary Methods Retrieving a complete string from the LZW tree therefore involves following the pointers in the parent fields. This is equivalent to moving up the tree, which is why the hash function is no longer needed. Example: The previous example describes the 15 hashing steps in the encoding of string alfeatsalfalfa. The last step sets array location 275 to (276,274,a) and writes 275 (a pointer to location 275) on the compressed stream. When this stream is read by the decoder, pointer 275 is the last item input and processed by the decoder. The decoder finds symbol a in the symbol field of location 275 (indicating that the string stored at 275 ends with an a) and a pointer to location 276 in the parent field. The decoder then examines location 276 where it finds symbol f and parent pointer 278. In location 278 the decoder finds symbol l and a pointer to 97. Finally, in location 97 the decoder finds symbol a and a null pointer. The (reversed) string is therefore afla. There is no need for the decoder to do any hashing or to use the index fields. The last point to discuss is string reversal. Two commonly-used approaches are outlined here: 1. Use a stack. A stack is a common data structure in modern computers. It is an array in memory that is accessed at one end only. At any time, the item that was last pushed into the stack will be the first one to be popped out (last-in-first-out, or LIFO). Symbols retrieved from the dictionary are pushed into the stack. When the last one has been retrieved and pushed, the stack is popped, symbol by symbol, into variable J. When the stack is empty, the entire string has been reversed. This is a common way to reverse a string. 2. Retrieve symbols from the dictionary and concatenate them into J from right to left. When done, the string will be stored in J in the right order. Variable J must be long enough to accommodate the longest possible string, but then it has to be long enough even when a stack is used.  Exercise 6.14: What is the longest string that can be retrieved from the LZW dictionary during decoding? (A reminder. The troublesome issue of software patents and licenses is treated in Section 6.34.) 6.13.3 LZW in Practice The publication of the LZW algorithm, in 1984, has strongly affected the data compression community and has influenced many people to come up with implementations and variants of this method. Some of the most important LZW variants and spin-offs are described here. 6.13.4 Differencing The idea of differencing, or relative encoding, has already been mentioned in Section 1.3.1. This idea turns out to be useful in LZW image compression, since most adjacent pixels don’t differ by much. It is possible to implement an LZW encoder that computes the value of a pixel relative to its predecessor and then encodes this difference. The decoder should, of course, be compatible and should compute the absolute value of a pixel after decoding its relative value. 6.14 UNIX Compression (LZC) 375 6.13.5 LZW Variants A word-based LZW variant is described in Section 11.6.2. LZW is an adaptive data compression method, but it is slow to adapt to its input, since strings in the dictionary become only one character longer at a time. Exercise 6.11 shows that a string of one million a’s (which, of course, is highly redundant) produces dictionary phrases, the longest of which contains only 1,414 a’s. The LZMW method, Section 6.15, is a variant of LZW that overcomes this problem. Its main principle is this: Instead of adding I plus one character of the next phrase to the dictionary, add I plus the entire next phrase to the dictionary. The LZAP method, Section 6.16, is yet another variant based on this idea: Instead of just concatenating the last two phrases and placing the result in the dictionary, place all prefixes of the concatenation in the dictionary. More specifically, if S and T are the last two matches, add St to the dictionary for every nonempty prefix t of T, including T itself. Table 6.27 summarizes the principles of LZW, LZMW, and LZAP and shows how they naturally suggest another variant, LZY. Increment Add a string to the dictionary string by per phrase per input char. One character: LZW LZY Several chars: LZMW LZAP Table 6.27: LZW and Three Variants. LZW adds one dictionary string per phrase and increments strings by one symbol at a time. LZMW adds one dictionary string per phrase and increments strings by several symbols at a time. LZAP adds one dictionary string per input symbol and increments strings by several symbols at a time. LZY, Section 6.18, fits the fourth cell of Table 6.27. It is a method that adds one dictionary string per input symbol and increments strings by one symbol at a time. 6.14 UNIX Compression (LZC) In the vast UNIX world, compress used to be the most common compression utility (although GNU gzip has become more popular because it is free from patent claims, is faster, and provides superior compression). This utility (also called LZC) uses LZW with a growing dictionary. It starts with a small dictionary of just 29 = 512 entries (with the first 256 of them already filled up). While this dictionary is being used, 9- bit pointers are written onto the output stream. When the original dictionary fills up, its size is doubled, to 1024 entries, and 10-bit pointers are used from this point. This process continues until the pointer size reaches a maximum set by the user (it can be set to between 9 and 16 bits, with 16 as the default value). When the largest allowed dictionary fills up, the program continues without changing the dictionary (which then becomes static), but with monitoring the compression ratio. If the ratio falls below a 376 6. Dictionary Methods predefined threshold, the dictionary is deleted, and a new 512-entry dictionary is started. This way, the dictionary never gets “too out of date.” Decoding is done by the uncompress command, which implements the LZC decoder. Its main task is to maintain the dictionary in the same way as the encoder. Two improvements to LZC, proposed by [Horspool 91], are listed below: 1. Encode the dictionary pointers with the phased-in binary codes of Section 5.3.1. Thus if the dictionary size is 29 = 512 entries, pointers can be encoded in either 8 or 9 bits. 2. Determine all the impossible strings at any point. Suppose that the current string in the look-ahead buffer is abcd... and the dictionary contains strings abc and abca but not abcd. The encoder will output, in this case, the pointer to abc and will start encoding a new string starting with d. The point is that after decoding abc, the decoder knows that the next string cannot start with an a (if it did, an abca would have been encoded, instead of abc). In general, if S is the current string, then the next string cannot start with any symbol x that satisfies “Sx is in the dictionary.” This knowledge can be used by both the encoder and decoder to reduce redundancy even further. When a pointer to a string should be output, it should be coded, and the method of assigning the code should eliminate all the strings that are known to be impossible at that point. This may result in a somewhat shorter code but is probably too complex to justify its use in practice. 6.14.1 LZT LZT is an extension, proposed by [Tischer 87], of UNIX compress/LZC. The major innovation of LZT is the way it handles a full dictionary. Recall that LZC treats a full dictionary in a simple way. It continues without updating the dictionary until the compression ratio drops below a preset threshold, and then it deletes the dictionary and starts afresh with a new dictionary. In contrast, LZT maintains, in addition to the dictionary (which is structured as a hash table, as described in Section 6.13.2), a linked list of phrases sorted by the number of times a phrase is used. When the dictionary fills up, the LZT algorithm identifies the least-recently-used (LRU) phrase in the list and deletes it from both the dictionary and the list. Maintaining the list requires several operations per phrase, which is why LZT is slower than LZW or LZC, but the fact that the dictionary is not completely deleted results in better compression performance (each time LZC deletes its dictionary, its compression performance drops significantly for a while). There is some similarity between the simple LZ77 algorithm and the LZT encoder. In LZ77, the search buffer is shifted, which deletes the leftmost phrase regardless of how useful it is (i.e., how often it has appeared in the input). In contrast, the LRU approach of LZT deletes the least useful phrase, the one that has been used the least since its insertion into the dictionary. In both cases, we can visualize a window of phrases. In LZ77 the window is sorted by the order in which phrases have been input, while in LZT the window is sorted by the usefulness of the phrases. Thus, LZT adapts better to varying statistics in the input data. Another innovation of LZT is its use of phased-in codes (Section 2.9). It starts with a dictionary of 512 entries, of which locations 0 through 255 are reserved for the 256 8-bit bytes (those are never deleted) and code 256 is reserved as an escape code. These 257 codes are encoded by 255 8-bit codes and two 9-bit codes. Once the first 6.15 LZMW 377 input phrase has been identified and added to the dictionary, it (the dictionary) has 258 codes and they are encoded by 254 8-bit codes and four 9-bit codes. The use of phased-in codes is especially useful at the beginning, when the dictionary contains only a few useful phrases. As encoding progresses and the dictionary fills up, these codes contribute less and less to the overall performance of LZT. The LZT dictionary is maintained as a hash table where each entry contains a phrase. Recall that in LZW (and also in LZC), if a phrase sI is in the dictionary (where s is a string and I is a single character), then s is also in the dictionary. This is why each dictionary entry in LZT has three fields (1) an output code (the code that has been assigned to a string s), (2) the (ASCII) code of an extension character I, and (3) a pointer to the list. When an entry is deleted, it is marked as free and the hash table is reorganized according to algorithm R in volume 3, Section 6.4 of [Knuth 73]. Another difference between LZC and LZT is the hash function. In LZC, the hash function takes an output code, shifts it one position to the left, and exclusive-ors it with the bits of the extension character. The hash function of LZT, on the other hand, reverses the bits of the output code before shifting and exclusive-oring it, which leads to fewer collisions and an almost perfect hash performance. 6.15 LZMW This LZW variant, developed by V. Miller and M. Wegman [Miller and Wegman 85], is based on two principles: 1. When the dictionary becomes full, the least-recently-used dictionary phrase is deleted. There are several ways to select this phrase, and the developers suggest that any reasonable way of doing so would work. One possibility is to identify all the dictionary phrases S for which there are no phrases Sa (nothing has been appended to S, implying that S hasn’t been used since it was placed in the dictionary) and delete the oldest of them. An auxiliary data structure has to be constructed and maintained in this case, pointing to dictionary phrases according to their age (the first pointer always points to the oldest phrase). The first 256 dictionary phrases should never be deleted. 2. Each phrase added to the dictionary is a concatenation of two strings, the previous match (S’ below) and the current one (S). This is in contrast to LZW, where each phrase added is the concatenation of the current match and the first symbol of the next match. The pseudo-code algorithm illustrates this: Initialize Dict to all the symbols of alphabet A; i:=1; S’:=null; while i <= input size k:=longest match of Input[i] to Dict; Output(k); S:=Phrase k of Dict; i:=i+length(S); If phrase S’S is not in Dict, append it to Dict; S’:=S; endwhile; 378 6. Dictionary Methods By adding the concatenation S’S to the LZMW dictionary, dictionary phrases can grow by more than one symbol at a time. This means that LZMW dictionary phrases are more “natural” units of the input (for example, if the input is text in a natural language, dictionary phrases will tend to be complete words or even several words in that language). This, in turn, implies that the LZMW dictionary generally adapts to the input faster than the LZW dictionary. Table 6.28 illustrates the LZMW method by applying it to the string sirsideastmaneasilyteasesseasickseals. LZMW adapts to its input faster than LZW but has the following three disadvantages: 1. The dictionary data structure cannot be the simple LZW trie, because not every prefix of a dictionary phrase is included in the dictionary. This means that the onesymbol- at-a-time search method used in LZW will not work. Instead, when a phrase S is added to the LZMW dictionary, every prefix of S must be added to the data structure, and every node in the data structure must have a tag indicating whether the node is in the dictionary or not. 2. Finding the longest string may require backtracking. If the dictionary contains aaaa and aaaaaaaa, we have to reach the eighth symbol of phrase aaaaaaab to realize that we have to choose the shorter phrase. This implies that dictionary searches in LZMW are slower than in LZW. This problem does not apply to the LZMW decoder. 3. A phrase may be added to the dictionary twice. This again complicates the choice of data structure for the dictionary.  Exercise 6.15: Use the LZMW method to compress the string swiss miss.  Exercise 6.16: Compress the string yabbadabbadabbadoo using LZMW. 6.16 LZAP LZAP is an extension of LZMW. The “AP” stands for “All Prefixes” [Storer 88]. LZAP adapts to its input fast, like LZMW, but eliminates the need for backtracking, a feature that makes it faster than LZMW. The principle is this: Instead of adding the concatenation S’S of the last two phrases to the dictionary, add all the strings S’t where t is a prefix of S (including S itself). Thus if S = a and S = bcd, add phrases ab,abc, and abcd to the LZAP dictionary. Table 6.29 shows the matches and the phrases added to the dictionary for yabbadabbadabbadoo. In step 7 the encoder concatenates d to the two prefixes of ab and adds the two phrases da and dab to the dictionary. In step 9 it concatenates ba to the three prefixes of dab and adds the resulting three phrases bad, bada, and badab to the dictionary. LZAP adds more phrases to its dictionary than does LZMW, so it takes more bits to represent the position of a phrase. At the same time, LZAP provides a bigger selection of dictionary phrases as matches for the input string, so it ends up compressing slightly better than LZMW while being faster (because of the simpler dictionary data structure, which eliminates the need for backtracking). This kind of trade-off is common in computer algorithms. 6.16 LZAP 379 Out- Add to Step Input put S dict. S’ sir sid eastman easily teases sea sick seals 1 s 115 s — — 2 i 105 i 256-si s 3 r 114 r 257-ir i 4 - 32  258-r r 5 si 256 si 259-si  6 d 100 d 260-sid si 7 - 32  261-d d 8 e 101 e 262-e  9 a 97 a 263-ea e 10 s 115 s 264-as a 11 t 117 t 265-st s 12 m 109 m 266-tm t 13 a 97 a 267-ma m 14 n 110 n 268-an a 15 -e 262 e 269-ne n 16 as 264 as 270-eas e 17 i 105 i 271-asi as 18 l 108 l 272-il i 19 y 121 y 273-ly l 20 - 32  274-y y 21 t 117 t 275-t  22 ea 263 ea 276-tea t 23 s 115 s 277-eas ea 24 e 101 e 278-se s 25 s 115 s 279-es e 26 - 32  280-s s 27 se 278 se 281-se  28 a 97 a 282-sea se 29 -si 259 si 283-asi a 30 c 99 c 284-sic si 31 k 107 k 285-ck c 32 -se 281 se 286-kse k 33 a 97 a 287-sea se 34 l 108 l 288-al a 35 s 115 s 289-ls l Table 6.28: LZMW Example. 380 6. Dictionary Methods Step Input Match Add to dictionary yabbadabbadabbadoo 1y y— 2a a256-ya 3 b b257-ab 4 b b258-bb 5 a a259-ba 6 d d260-ad 7 ab ab 261-da, 262-dab 8 ba ba 263-abb, 264-abba 9 dab dab 265-bad, 266-bada, 267-badab 10 bad bad 268-dabb, 269-dabba, 270-dabbad 11 o o 271-bado 12 o o 272-oo Table 6.29: LZAP Example. 6.17 LZJ LZJ is an interesting LZ variant, originated by [Jakobsson 85]. It stores in its dictionary, which can be viewed either as a multiway tree or as a forest, every phrase found in the input. If a phrase is found n times in the input, only one copy is stored in the dictionary. Such behavior tends to fill the dictionary up very quickly, so LZJ limits the length of phrases to a preset parameter h. The developer of the method recommends h values of around 6. Thus, if h = 4, the alphabet includes the letters and the blank space, and the input is Onceuponatime..., then the dictionary will start with all the symbols of the alphabet and the following strings will gradually be inserted in it: on, onc, nc, ce, nce, Once, nce, ceu, eup, upo, and so on. (The value of h is critical. Small values of h cause the LZJ encoder to store only short phrases in the dictionary. This degrades the compression because every code in the compressed file will correspond to only a short string. Large values of h also degrade compression because only a few long strings occur often in the input.) An indirect result of this principle and of the way LZJ operates is that LZJ encoding is simpler than its decoding. Thus, LZJ is an asymmetric compression method and is unusual in this respect because in other asymmetric methods decoding is simpler and faster than encoding. Example. Given the alphabet a, b, and c, the value h = 4, and the input data babbaaacba, the left side of Figure 6.30 lists the 24 phrases whose length is h or shorter, together with their dictionary addresses. The right side shows the complete 3-way dictionary tree (as a forest of three trees) with all 24 phrases. Each node in the dictionary has four fields, the character itself, a pointer (shown as a solid arrow) to its leftmost child, a pointer (horizontal, in dashed) to its right sibling, if any, and a frequency count that is discussed below. If a node is a rightmost child, the sibling pointer is replaced by a pointer to the parent (dashed, pointing up). The dictionary’s size H is another parameter of LZJ and is preset in each implementation (in the mid 1980s, the developer recommended H . 213, but current imple- 6.17 LZJ 381 a 0 bb 9 aac 17 b 1 bba 10 aacb 18 c 2 bbba 11 ac 19 ba 3 baa 12 acb 20 bab 4 baaa 13 acba 21 babb 5 aa 14 cb 22 ab 6 aaa 15 cba 23 abb 7 aaac 16 abba 8                         Figure 6.30: An LZJ Dictionary. mentations can afford much larger values). Each input phrase found by the encoder in the dictionary is compressed by writing the dictionary address of the last character of the phrase on the output. This address is a number between 0 and H - 1, written in log2 H bits. In order to understand the operation of the encoder, we assume h = 4 and the input string Onceuponatime.... The encoder initializes the dictionary to the characters of the alphabet. At a general point in the encoding it reads the next input string, say ponatime..., selects the dictionary tree for p, and tries to match as many characters as possible. If pon is found in the tree, but pon isn’t, the encoder outputs the dictionary address of the n of pon and updates the dictionary by inserting in it the strings eup, upo, and upon. They are placed in the trees whose roots are e, , and u, respectively. As each string is inserted, the encoder increments the counts of all the nodes that are on the path from the root to the leaf node of the string. When the dictionary fills up, the encoder prunes the trees by deleting any nodes with a count of 1 (where instead of 1, an implementation can opt for a user-defined parameter). The roots of the trees correspond to the characters of the alphabet and should never be removed. The decoder can mimic this operation in lockstep with the encoder. Pruning the trees results in free nodes, but consecutive prunings result in fewer and fewer free nodes, as illustrated by Figure 6.31. The figure shows that with H = 1,024, many dictionary prunings are needed for the first 4,000 input characters. The developer of the method also shows that doubling the dictionary size to 2,048 nodes results in only three dictionary prunings for the same input data. Eventually, pruning becomes ineffective and LZJ continues (nonadaptively) without it. This behavior suggests that a large dictionary may not require any prunings at all and may therefore speed up LZJ (both encoding and decoding) considerably. LZJ decoding is more time consuming than its encoding, because the decoder has to follow various pointers in a dictionary tree. This is best illustrated by an example (refer to Figure 6.30). Suppose that the decoder reads 17 from the compressed stream. Location 17 contains the character c, the last character in the phrase being decoded. The decoder has to follow the back pointers (the dashed pointers in the figure) until it arrives at the root of the tree. It then follows the forward pointers (both dashed horizontal and solid), collecting characters as it goes along, until it arrives back at location 17. This is 382 6. Dictionary Methods       Figure 6.31: Effect of Pruning the LZJ Dictionary. how the decoder constructs phrase aac. Once the phrase is decoded, the decoder uses it to update the dictionary by inserting phrases in lockstep with the encoder. The last interesting feature of LZJ is the format of the compressed stream. If the dictionary size H is fixed, then each item in the compressed stream is a fixed-size, log2 H bits pointer to the dictionary. In principle, however, the dictionary can be allowed to grow as needed and the algorithm may use one of the many variable-length codes for the integers to encode the dictionary pointers. It turns out that this approach, which initially seems promising, is in fact worse than having fixed-length pointers. The reason is that each variable-length code is ideal for a certain distribution of the integers in the data, but in the case of LZJ nothing is known about this distribution. Initially, the dictionary is small, so all the pointers to it are small integers. When the dictionary grows, especially after it has been pruned, pointers to it may be small or large with the same probability. Given a set of n integers with a uniform distribution, the length of a fixed-length code for them is log2 n, but the average length of any variable-length code is generally longer (see Figure 3.11). A good compromise is the use of phased-in codes (Section 2.9). If H = 1,000, the length of a fixed-size code is 10 bits, but only 1,000 out of the 1,024 possible 10-bit combinations are used. With phased-in codes, most codewords would be 10 bits long, but some codewords would be only nine bits long. The excellent book Text Compression [Bell et al. 90], refers to this compromise (on page 230) as LZJ. “Heavens, Andrew!” said his wife; “what is a rustler?” It was not in any dictionary, and current translations of it were inconsistent. A man at Hoosic Falls said that he had passed through Cheyenne, and heard the term applied in a complimentary way to people who were alive and pushing. Another man had always supposed it meant some kind of horse. But the most alarming version of all was that a rustler was a cattle thief. —Owen Wister, The Virginian—A Horseman of the Plains 6.18 LZY 383 6.18 LZY The LZY method is due to Dan Bernstein. The Y stands for Yabba, which came from the input string originally used to test the algorithm. The LZY dictionary is initialized to all the single symbols of the alphabet. For every symbol C in the input stream, the decoder looks for the longest string P that precedes C and is already included in the dictionary. If the string PC is not in the dictionary, it is added to it as a new phrase. As an example, the input yabbadabbadabbadoo causes the phrases ya, ab, bb, ba, ad, da, abb, bba, ada, dab, abba, bbad, bado, ado, and oo to be added to the dictionary. While encoding the input, the encoder keeps track of the list of matches-so-far L. Initially, L is empty. If C is the current input symbol, the encoder (before adding anything to the dictionary) checks, for every string M in L, whether string MC is in the dictionary. If it is, then MC becomes a new match-so-far and is added to L. Otherwise, the encoder outputs the number of L (its position in the dictionary) and adds C, as a new match-so-far, to L. Here is a pseudo-code algorithm for constructing the LZY dictionary. The authors’ personal experience suggests that implementing such an algorithm in a real programming language results in a deeper understanding of its operation. Start with a dictionary containing all the symbols of the alphabet, each mapped to a unique integer. M:=empty string. Repeat Append the next symbol C of the input stream to M. If M is not in the dictionary, add it to the dictionary, delete the first character of M, and repeat this step. Until end-of-input. The output of LZY is not synchronized with the dictionary additions. Also, the encoder must be careful not to have the longest output match overlap itself. Because of this, the dictionary should consist of two parts, S and T, where only the former is used for the output. The algorithm is the following: Start with S mapping each single character to a unique integer; set T empty; M empty; and O empty. Repeat Input the next symbol C. If OC is in S, set O:=OC; otherwise output S(O), set O:=C, add T to S, and remove everything from T. While MC is not in S or T, add MC to T (mapping to the next available integer), and chop off the first character of M. After M is short enough so that MC is in the dict., set M:=MC. Until end-of-input. Output S(O) and quit. The decoder reads the compressed stream. It uses each code to find a phrase in the dictionary, it outputs the phrase as a string, then uses each symbol of the string to add a new phrase to the dictionary in the same way the encoder does. Here are the decoding steps: 384 6. Dictionary Methods Start with a dictionary containing all the symbols of the alphabet, each mapped to a unique integer. M:=empty string. Repeat Read D(O) from the input and take the inverse under D to find O. As long as O is not the empty string, find the first character C of O, and update (D,M) as above. Also output C and chop it off from the front of O. Until end-of-input. Notice that encoding requires two fast operations on strings in the dictionary: (1) testing whether string SC is in the dictionary if S’s position is known and (2) finding S’s position given CS’s position. Decoding requires the same operations plus fast searching to find the first character of a string when its position in the dictionary is given. Table 6.32 illustrates LZY for the input string abcabcabcabcabcabcabcx’. It shows the phrases added to the dictionary at each step, as well as the list of current matches. The encoder starts with no matches. When it inputs a symbol, it appends it to each match-so-far; any results that are already in the dictionary become the new matches-sofar (the symbol itself becomes another match). Any results that are not in the dictionary are deleted from the list and added to the dictionary. Before reading the fifth c, for example, the matches-so-far are bcab, cab, ab, and b. The encoder appends c to each match. bcabc doesn’t match, so the encoder adds it to the dictionary. The rest are still in the dictionary, so the new list of matches-so-far is cabc, abc, bc, and c. When the x is input, the current list of matches-so-far is abcabc, bcabc, cabc, abc, bc, and c. None of abcabcx, bcabcx, cabcx, abcx, bcx, or cx are in the dictionary, so they are all added to it, and the list of matches-so-far is reduced to just a single x. Airman stops coed Anagram of “data compression” 6.19 LZP LZP is an LZ77 variant developed by Charles Bloom [Bloom 96] (the P stands for “prediction”). It is based on the principle of context prediction, which says, “if a certain string abcde has appeared in the input stream in the past and was followed by fg..., then when abcde appears again in the input stream, there is a good chance that it will be followed by the same fg....” Section 6.35 should be consulted for the relation between dictionary-based and prediction algorithms. Figure 6.33 (part I) shows an LZ77 sliding buffer with fgh... as the current symbols (this string is denoted by S) in the look-ahead buffer, immediately preceded by abcde in the search buffer. The string abcde is called the context of fgh... and is denoted by C. In general, the context of a string S is the N-symbol string C immediately to the left of S. A context can be of any length N, and variants of LZP, discussed in Sections 6.19.3 and 6.19.4, use different values of N. The algorithm passes the context through a hash 6.19 LZP 385 Step Input Add to dict. Current matches abcabcabcabcabcabcabcx 1 a — a 2 b 256-ab b 3 c 257-bc c 4 a 258-ca a 5 b — ab, b 6 c 259-abc bc, c 7 a 260-bca ca, a 8 b 261-cab ab, b 9 c — abc, bc, c 10 a 262-abca bca, ca, a 11 b 263-bcab cab, ab, b 12 c 264-cabc abc, bc, c 13 a — abca, bca, ca, a 14 b 265-abcab bcab, cab, ab, b 15 c 266-bcabc cabc, abc, bc, c 16 a 267-cabca abca, bca, ca, a 17 b — abcab, bcab, cab, ab, b 18 c 268-abcabc bcabc, cabc, abc, bc, c 19 a 269-bcabca cabca, abca, bca, ca, a 20 b 270-cabcab abcab, bcab, cab, ab, b 21 c — abcabc, bcabc, cabc, abc, bc, c 22 x 271-abcabcx x 23 272-bcabcx 24 273-cabcx 25 274-abcx 26 275-bcx 27 276-cx Table 6.32: LZY Example. ...abcdefgi... Search Buffer Look-Ahead Buffer Index Table P ...abcdefgh... Hash Function H Figure 6.33: The Principle of LZP: Part I. function and uses the result H as an index to a table of pointers called the index table. The index table contains pointers to various symbols in the search buffer. Index H is 386 6. Dictionary Methods used to select a pointer P. In a typical case, P points to a previously seen string whose context is also abcde (see below for atypical cases). The algorithm then performs the following steps: Step 1: It saves P and replaces it in the index table with a fresh pointer Q pointing to fgh... in the look-ahead buffer (Figure 6.33 Part II). An integer variable L is set to zero. It is used later to indicate the match length. Search Buffer Look-Ahead Buffer Index Table P Q ...abcdefgi... ...abcdefgh... Figure 6.33: The Principle of LZP: Part II. Step 2: If P is not a null pointer, the algorithm follows it and compares the string pointed at by P (string fgi... in the search buffer) to the string “fgh...” in the look-ahead buffer. Only two symbols match in our example, so the match length, L, is set to 2. Step 3: If L = 0 (no symbols have been matched), the buffer is slid to the right (or, equivalently, the input is shifted to the left) one position and the first symbol of string S (the f) is written on the compressed stream as a raw ASCII code (a literal). Step 4: IfL > 0 (L symbols have been matched), the buffer is slid to the right L positions and the value of L is written on the compressed stream (after being suitably encoded). In our example the single encoded value L = 2 is written on the compressed stream instead of the two symbols fg, and it is this step that produces compression. Clearly, the larger the value of L, the better the compression. Large values of L result when an Nsymbol context C in the input stream is followed by the same long string S as a previous occurrence of C. This may happen when the input stream features high redundancy. In a random input stream each occurrence of the same context C is likely to be followed by another S, leading to L = 0 and therefore to no compression. An “average” input stream results in more literals than L values being written on the output stream (see also Exercise 6.18). The decoder inputs the compressed stream item by item and creates the decompressed output in a buffer B. The steps are: Step 1: Input the next item I from the compressed stream. Step 2: If I is a raw ASCII code (a literal), it is appended to buffer B, and the data in B is shifted to the left one position. Step 3: If I is an encoded match length, it is decoded, to obtain L. The present context C (the rightmost N symbols in B) is hashed to an index H, which is used to select a pointer P from the index table. The decoder copies the string of L symbols starting at B[P] and appends it to the right end of B. It also shifts the data in B to the left L positions and replaces P in the index table with a fresh pointer, to keep in lockstep with the encoder. 6.19 LZP 387 Two points remain to be discussed before we are ready to look at a detailed example. 1. When the encoder starts, it places the first N symbols of the input stream in the search buffer, to become the first context. It then writes these symbols, as literals, on the compressed stream. This is the only special step needed to start the compression. The decoder starts by reading the first N items off the compressed stream (they should be literals), and placing them at the rightmost end of buffer B, to serve as the first context. 2. It has been mentioned before that in the typical case, P points to a previously-seen string whose context is identical to the present context C. In an atypical case, P may be pointing to a string whose context is different. The algorithm, however, does not check the context and always behaves in the same way. It simply tries to match as many symbols as possible. At worst, zero symbols will match, leading to one literal written on the compressed stream. 6.19.1 Example The input stream xyabcabcabxy is used to illustrate the operation of the LZP encoder. To simplify the example, we use N = 2. xyabcabcabxy (a) H 0 1 7 xyabcabcabxy (b) 0 1 7 xyabcabcabxy H (c) 0 1 xyabcabcabxy (d) 0 1 Hash ya Hash ya Figure 6.34: LZP Compression of xyabcabcabxy: Part I. 1. To start the operation, the encoder shifts the first two symbols xy to the search buffer and outputs them as literals. It also initializes all locations of the index table to the null pointer. 2. The current symbol is a (the first a), and the context is xy. It is hashed to, say, 5, but location 5 of the index table contains a null pointer, so P is null (Figure 6.34a). Location 5 is set to point to the first a (Figure 6.34b), which is then output as a literal. The data in the encoder’s buffer is shifted to the left. 3. The current symbol is the first b, and the context is ya. It is hashed to, say, 7, but location 7 of the index table contains a null pointer, so P is null (Figure 6.34c). Location 388 6. Dictionary Methods 7 is set to point to the first b (Figure 6.34d), which is then output as a literal. The data in the encoder’s buffer is shifted to the left. 4. The current symbol is the first c, and the context is ab. It is hashed to, say, 2, but location 2 of the index table contains a null pointer, so P is null (Figure 6.34e). Location 2 is set to point to the first c (Figure 6.34f), which is then output as a literal. The data in the encoder’s buffer is shifted to the left. 5. The same happens two more times, writing the literals a and b on the compressed stream. The current symbol is now (the second) c, and the context is ab. This context is hashed, as in step 4, to 2, so P points to “cabc...”. Location 2 is set to point to the current symbol (Figure 6.34g), and the encoder tries to match strings cabcabxy and cabxy. The resulting match length is L = 3. The number 3 is written, encoded on the output, and the data is shifted three positions to the left. 6. The current symbol is the second x, and the context is ab. It is hashed to 2, but location 2 of the index table points to the second c (Figure 6.34h). Location 2 is set to point to the current symbol, and the encoder tries to match strings cabxy and xy. The resulting match length is, of course, L = 0, so the encoder writes x on the compressed stream as a literal and shifts the data one position. 7. The current symbol is the second y, and the context is bx. It is hashed to, say, 7. This is a hash collision, since context ya was hashed to 7 in step 3, but the algorithm does not check for collisions. It continues as usual. Location 7 of the index table points to the first b (or rather to the string bcabcabxy). It is set to point to the current symbol, and the encoder tries to match strings bcabcabxy and y, resulting in L = 0. The encoder writes y on the compressed stream as a literal and shifts the data one position. 8. The current symbol is the end-of-data, so the algorithm terminates. P xyabcabcabxy (g) xyabcabcabxy (h) old P new P H xyabcabcabxy (e) 0 1 H xyabcabcabxy (f) 0 1 Hash ab Hash ab Figure 6.34 (Continued). LZP Compression of xyabcabcabxy: Part II.  Exercise 6.17: Write the LZP encoding steps for the input string xyaaaa.... Structures are the weapons of the mathematician. —Nicholas Bourbaki 6.19 LZP 389 6.19.2 Practical Considerations Shifting the data in the buffer would require updating all the pointers in the index table. An efficient implementation should therefore adopt a better solution. Two approaches are described below, but other ones may also work. 1. Reserve a buffer as large as possible, and compress the input stream in blocks. Each block is input into the buffer and is never shifted. Instead, a pointer is moved from left to right, to point at any time to the current symbol in the buffer. When the entire buffer has been encoded, the buffer is filled up again with fresh input. This approach simplifies the program but has the disadvantage that the data at the end of a block cannot be used to predict anything for the next block. Each block is encoded independently of the other ones, leading to poorer compression. 2. Reserve a large buffer, and use it as a circular queue (Section 6.3.1). The data itself does not have to be shifted, but after encoding the current symbol the data is effectively shifted by updating the start and end pointers, and a new symbol is input and stored in the buffer. The algorithm is somewhat more complicated, but this approach has the advantage that the entire input is encoded as one stream. Every symbol benefits from the D symbols preceding it (where D is the total length of the buffer). Imagine a pointer P in the index table pointing to some symbol X in the buffer. When the movement of the two pointers in the circular queue leaves X outside the queue, some new symbol Y will be input into the position occupied by X, and P will now be pointing to Y. When P is next selected by the hashing function and is used to match strings, the match will likely result in L = 0. However, the algorithm always replaces the pointer that it uses, so such a case should not degrade the algorithm’s performance significantly. 6.19.3 LZP1 and LZP2 There are currently four versions of LZP, called LZP1 through LZP4. This section discusses the details of the first two. The context used by LZP1 is of order 3, i.e., it is the three bytes preceding the current one. The hash function produces a 12-bit index H and is best described by the following C code: H=((C>>11)^C)&0xFFF. Since H is 12 bits, the index table should be 212 = 4,096 entries long. Each entry is two bytes (16 bits), but only 14 of the 16 bits are used. A pointer P selected in the index table thus points to a buffer of size 214 = 16K. The LZP1 encoder creates a compressed stream with literals and L values mixed together. Each item must therefore be preceded by a flag indicating its nature. Since only two flags are needed, the simplest choice would be 1-bit flags. However, we have already mentioned that an “average” input stream results in more literals than L values, so it makes sense to assign a short flag (less than one bit) to indicate a literal, and a long flag (a wee bit longer than one bit) to indicate a length. The scheme used by LZP1 uses 1 to indicate two literals, 01 to indicate a literal followed by a match length, and 00 to indicate a match length.  Exercise 6.18: Let T indicate the probability of a literal in the compressed stream. For what value of T does the above scheme produce flags that are 1-bit long on average? 390 6. Dictionary Methods A literal is written on the compressed stream as an 8-bit ASCII code. Match lengths are encoded according to Table 6.35. Initially, the codes are 2 bits. When these are all used up, 3 bits are added, for a total of 5 bits, where the first 2 bits are 1’s. When these are also all used, 5 bits are added, to produce 10-bit codes where the first 5 bits are 1’s. From then on another group of 8 bits is added to the code whenever all the old codes have been used up. Notice how the pattern of all 1’s is never used as a code and is reserved to indicate longer and longer codes. Notice also that a unary code or a general unary code (Section 3.1) might have been a better choice. Length Code Length Code 1 00 11 11|111|00000 2 01 12 11|111|00001 3 10 ... 4 11| 000 41 11|111|11110 5 11| 001 42 11|111|11111|00000000 6 11| 010 ... ... 296 11|111|11111|11111110 10 11|110 297 11|111|11111|11111111|00000000 Table 6.35: Codes Used by LZP1 and LZP2 for Match Lengths. The compressed stream consists of a mixture of literals (bytes with ASCII codes) and control bytes containing flags and encoded lengths. This is illustrated by the output of the example of Section 6.19.1. The input of this example is the string xyabcabcabxy, and the output items are x, y, a, b, c, a, b, 3, x, and y. The actual output stream consists of the single control byte 111 01|10 1 followed by nine bytes with the ASCII codes of x, y, a, b, c, a, b, x, and y.  Exercise 6.19: Explain the content of the control byte 111 01|10 1. Another example of a compressed stream is the three literals x, y, and a followed by the four match lengths 12, 12, 12, and 10. We first prepare the flags 1 (x, y) 01 (a, 12) 00 (12) 00 (12) 00 (12) 00 (10), then substitute the codes of 12 and 10, 1xy01a11|111|0000100|11|111|0000100|11|111|0000100|11|111|0000100|11|110, and finally group together the bits that make up the control bytes. The result is 10111111 x, y, a, 00001001 11110000 10011111 00001001 11110000 10011110. Notice that the first control byte is followed by the three literals. The last point to be mentioned is the case ...0 1yyyyyyy zzzzzzzz. The first control byte ends with the 0 of a pair 01, and the second byte starts with the 1 of the same pair. This indicates a literal followed by a match length. The match length is the yyy bits (at least some of them) in the second control byte. If the code of the match length is long, then the zzz bits or some of them may be part of the code. The literal is either the zzz byte or the byte following it. LZP2 is identical to LZP1 except that literals are coded using nonadaptive Huffman codes. Ideally, two passes should be used; the first one counting the frequency of 6.20 Repetition Finder 391 occurrence of the symbols in the input stream and the second pass doing the actual compression. In between the passes, the Huffman code table can be constructed. 6.19.4 LZP3 and LZP4 LZP3 is similar to both LZP1 and LZP2. It uses order-4 contexts and more sophisticated Huffman codes to encode both the match lengths and the literals. The LZP3 hash function is H=((C>>15)^C)&0xFFFF, so H is a 16-bit index, to be used with an index table of size 216 = 64 K. In addition to the pointers P, the index table contains also the contexts C. Thus if a context C is hashed to an index H, the encoder expects to find the same context C in location H of the index table. This is called context confirmation. If the encoder finds something else, or if it finds a null pointer, it sets P to null and hashes an order-3 context. If the order-3 context confirmation also fails, the algorithm hashes the order-2 context, and if that also fails, the algorithm sets P to null and writes a literal on the compressed stream. This method attempts to find the highest-order context seen so far. LZP4 uses order-5 contexts and a multistep lookup process. In step 1, the rightmost four bytes I of the context are hashed to create a 16-bit index H according to the following: H=((I>>15)^I)&0xFFFF. Then H is used as an index to the index table that has 64K entries, each corresponding to one value of H. Each entry points to the start of a list linking nodes that have the same hash value. Suppose that the contexts abcde, xbcde, and mnopq hash to the same index H = 13 (i.e., the hash function computes the same index 13 when applied to bcde and nopq) and we are looking for context xbcde. Location 13 of the index table would point to a list with nodes for these contexts (and possibly others that have also been hashed to 13). The list is traversed until a node is found with bcde. This node points to a second list linking a, x, and perhaps other symbols that precede bcde. The second list is traversed until a node with x is found. That node contains a pointer to the most recent occurrence of context xbcde in the search buffer. If a node with x is not found, a literal is written to the compressed stream. This complex lookup procedure is used by LZP4 because a 5-byte context does not fit comfortably in a single word in most current computers. 6.20 Repetition Finder All the dictionary-based methods described so far have one thing in common: they use a large memory buffer as a dictionary that holds fragments of text found so far. The dictionary is used to locate strings of symbols that repeat. The method described here is different. Instead of a dictionary it uses a fixed-size array of integers to find previous occurrences of strings of text. The array size equals the square of the alphabet size, so it is not very large. The method is due to Hidetoshi Yokoo [Yokoo 91], who elected not to call it LZHY but left it nameless. The reason a name of the form LZxx was not used is that the method does not employ a traditional Ziv-Lempel dictionary. The reason it was 392 6. Dictionary Methods left nameless is that it does not compress very well and should therefore be considered the first step in a new field of research rather than a mature, practical method. The method alternates between two modes, normal and repeat. It starts in the normal mode, where it inputs symbols and encodes them using adaptive Huffman. When it identifies a repeated string it switches to the “repeat” mode where it outputs an escape symbol, followed by the length of the repeated string. Assume that the input stream consists of symbols x1x2 . . . from an alphabet A. Both encoder and decoder maintain an array REP of dimensions |A|Χ|A| that is initialized to all zeros. For each input symbol xi, the encoder (and decoder) compute a value yi according to yi = i-REP[xi-1, xi], and then update REP[xi-1, xi] := i. The 13-symbol string xi: X A B C D E Y A B C D E Z i : 1 3 5 7 9 11 13 results in the following y values: i= 1 2 3 4 5 6 7 8 9 10 11 12 13 yi= 1 2 3 4 5 6 7 8 6 6 6 6 13 xi-1xi: XA AB BC CD DE EY YA AB BC CD DE EZ Table 6.36a shows the state of array REP after the eighth symbol has been input and encoded. Table 6.36b shows the state of REP after all 13 symbols have been input and encoded. A B C D E . . . X Y Z A 3 B 4 C 5 D 6 E 7 ... X 2 Y 8 Z A B C D E . . . X Y Z A 9 B 10 C 11 D 12 E 13 ... X 2 Y 8 Z Table 6.36: (a) REP at i = 8. (b) REP at i = 13. Perhaps a better way to explain the way y is calculated is by the expression yi =... 1, for i = 1, i, for i > 1 and first occurrence of xi-1xi, min(k), for i > 1 and xi-1xi identical to xi-k-1xi-k. This shows that y is either i or is the distance k between the current string xi-1xi and its most recent copy xi-k-1xi-k. However, recognizing a repetition of a string is done 6.20 Repetition Finder 393 by means of array REP and without using any dictionary (which is the main point of this method). When a string of length l repeats itself in the input, l consecutive identical values of y are generated, and this is the signal for the encoder to switch to the “repeat” mode. As long as consecutive different values of y are generated, the encoder stays in the “normal” mode, where it encodes xi in adaptive Huffman and outputs it. When the encoder senses yi+1 = yi, it outputs xi in the normal mode, and then enters the “repeat” mode. In the example above this happens for i = 9, so the string XABCDEYAB is output in the normal mode. Once in the “repeat” mode, the encoder inputs more symbols and calculates y values until it finds the first value that differs from yi. In our example this happens at i = 13, when the Z is input. The encoder compresses the string “CDE” (corresponding to i = 10, 11, 12) in the “repeat” mode by emitting an (encoded) escape symbol, followed by the (encoded) length of the repeated string (3 in our example). The encoder then switches back to the normal mode, where it saves the y value for Z as yi and inputs the next symbol. The escape symbol must be an extra symbol, one that’s not included in the alphabet A. Notice that only two y values, yi-1 and yi, need be saved at any time. Notice also that the method is not very efficient, since it senses the repeating string “too late” and encodes the first two repeating symbols in the normal mode. In our example only three of the five repeating symbols are encoded in the “repeat” mode. The decoder inputs and decodes the first nine symbols, decoding them into the string XABCDEYAB while updating array REP and calculating y values. When the escape symbol is input, i has the value 9 and yi has the value 6. The decoder inputs and decodes the length, 3, and now it has to figure out the repeated string of length 3 using just the data in array REP, not any of the previously decoded input. Since i = 9 and yi is the distance between this string and its copy, the decoder knows that the copy started at position i - yi = 9- 6 = 3 of the input. It scans REP, looking for a 3. It finds it at position REP[A,B], so it starts looking for a 4 in row B of REP. It finds it in REP[B,C], so the first symbol of the required string is C. Looking for a 5 in row C, the decoder finds it in REP[C,D], so the second symbol is D. Looking now for a 6 in row D, the decoder finds it in REP[D,E]. This is how a repeated string can be decoded without maintaining a dictionary. Both encoder and decoder store values of i in REP, so an entry of REP should be at least two bytes long. This way i can have values of up to 64K-1 . 65,500, so the input has to be encoded in blocks of size 64K. For an alphabet of 256 symbols, the size of REP should therefore be 256 Χ 256 Χ 2 = 128 Kbytes, not very large. For larger alphabets REP may have to be uncomfortably large. In the normal mode, symbols (including the escape) are encoded using adaptive Huffman (Section 5.3). In the repeat mode, lengths are encoded in a recursive prefix code denoted Qk(i), where k is a positive integer (see Section 2.2 for prefix codes). Assuming that i is an integer whose binary representation is 1., the prefix code Qk(i) of i is defined by Q0(i) = 1|.|0., Qk(i) = 0, i= 1, 1Qk-1(i - 1), i>1, 394 6. Dictionary Methods where |.| is the length of . and 1|.| is a string of |.| ones. Table 6.37 shows some of the proposed codes; however, any of the prefix codes of Section 2.2 can be used instead of the Qk(i) codes proposed here. i . Q0 (i) Q1(i) Q2(i) 1 null 0 0 0 2 0 100 10 10 3 1 101 1100 110 4 00 11000 1101 11100 5 01 11001 111000 11101 6 10 11010 111001 1111000 7 11 11011 111010 1111001 8 000 1110000 111011 1111010 9 001 1110001 11110000 1111011 Table 6.37: Proposed Prefix Code. The developer of this method, Hidetoshi Yokoo, indicates that compression performance is not very sensitive to the precise value of k, and he proposes k = 2 for best overall performance. As mentioned earlier, the method is not very efficient, which is why it should be considered the start of a new field of research where repeated strings are identified without the need for a large dictionary. 6.21 GIF Images GIF—the graphics interchange format—was developed by Compuserve Information Services in 1987 as an efficient, compressed graphics file format, which allows for images to be sent between different computers. The original version of GIF is known as GIF 87a. The current standard is GIF 89a and, at the time of writing, can be freely obtained as the file http://delcano.mit.edu/info/gif.txt. GIF is not a data compression method; it is a graphics file format that uses a variant of LZW to compress the graphics data (see [Blackstock 87]). This section reviews only the data compression aspects of GIF. In compressing data, GIF is very similar to compress and uses a dynamic, growing dictionary. It starts with the number of bits per pixel b as a parameter. For a monochromatic image, b = 2; for an image with 256 colors or shades of gray, b = 8. The dictionary starts with 2b+1 entries and is doubled in size each time it fills up, until it reaches a size of 212 = 4,096 entries, where it remains static. At such a point, the encoder monitors the compression ratio and may decide to discard the dictionary at any point and start with a new, empty one. When making this decision, the encoder emits the value 2b as the clear code, which is the sign for the decoder to discard its dictionary. The pointers, which get longer by 1 bit from one dictionary to the next, are accumulated and are output in blocks of 8-bit bytes. Each block is preceded by a header 6.22 RAR and WinRAR 395 that contains the block size (255 bytes maximum) and is terminated by a byte of eight zeros. The last block contains, just before the terminator, the eof value, which is 2b +1. An interesting feature is that the pointers are stored with their least significant bit on the left. Consider, for example, the following 3-bit pointers 3, 7, 4, 1, 6, 2, and 5. Their binary values are 011, 111, 100, 001, 110, 010, and 101, so they are packed in 3 bytes |10101001|11000011|11110...|. The GIF format is commonly used today by web browsers, but it is not an efficient image compressor. GIF scans the image row by row, so it discovers pixel correlations within a row, but not between rows. We can say that GIF compression is inefficient because GIF is one-dimensional while an image is two-dimensional. An illustrative example is the two simple images of Figure 7.4a,b (Section 7.3). Saving both in GIF89 has resulted in file sizes of 1053 and 1527 bytes, respectively. Most graphics file formats use some kind of compression. For more information on those files, see [Murray and vanRyper 94]. 6.22 RAR and WinRAR The popular RAR software is the creation of Eugene Roshal, who started it as his university doctoral dissertation. RAR is an acronym that stands for Roshal ARchive (or Roshal ARchiver). The current developer is Eugene’s brother Alexander Roshal. The following is a list of its most important features: RAR is currently available from [rarlab 06], as shareware, for Windows (WinRAR), Pocket PC, Macintosh OS X, Linux, DOS, and FreeBSD. WinRAR has a graphical user interface, whereas the other versions support only a command line interface. WinRAR provides complete support for RAR and ZIP archives and can decompress (but not compress) CAB, ARJ, LZH, TAR, GZ, ACE, UUE, BZ2, JAR, ISO, 7Z, and Z archives. In addition to compressing data, WinRAR can encrypt data with the advanced encryption standard (AES-128). WinRAR can compress files up to 8,589 billion Gb in size (approximately 9Χ1018 bytes). Files compressed in WinRAR can be self-extracting (SFX) and can come from different volumes. It is possible to combine many files, large and small, into a so-called “solid” archive, where they are compressed together. This may save 10–50% compared to compressing each file individually. The RAR software can optionally generate recovery records and recovery volumes that make it possible to reconstruct damaged archives. Redundant data, in the form of a Reed-Solomon error-correcting code, can optionally be added to the compressed file for increased reliability. 396 6. Dictionary Methods The user can specify the amount of redundancy (as a percentage of the original data size) to be built into the recovery records of a RAR archive. A large amount of redundancy makes the RAR file more resistant to data corruption, thereby allowing recovery from more serious damage. However, any redundant data reduces compression efficiency, which is why compression and reliability are opposite concepts. Authenticity information may be included for extra security. RAR software saves information on the last update and name of the archive. An intuitive wizard especially designed for novice users. The wizard makes it easy to use all of RAR’s features through a simple question and answer procedure. Excellent drag-and-drop features. The user can drag files from WinRAR to other programs, to the desktop, or to any folder. An archive dropped on WinRAR is immediately opened to display its files. A file dropped on an archive that’s open in WinRAR is immediately added to the archive. A group of files or folders dropped on WinRAR automatically becomes a new archive. RAR offers NTFS and Unicode support (see [ntfs 06] for NTFS). WinRAR is available in over 35 languages. These features, combined with excellent compression, good compression speed, an attractive graphical user interface, and backup facilities, make RAR one of the best overall compression tools currently available. The RAR algorithm is generally considered proprietary, and the following quote (from [donationcoder 06]) sheds some light on what this implies The fact that the RAR encoding algorithm is proprietary is an issue worth considering. It means that, unlike ZIP and 7z and almost all other compression algorithms, only the official WinRAR programs can create RAR archives (although other programs can decompress them). Some would argue that this is unhealthy for the software industry and that standardizing on an open format would be better for everyone in the long run. But for most users, these issues are academic; WinRAR offers great support for both ZIP and RAR formats. Proprietary A term that indicates that a party, or proprietor, exercises private ownership, control, or use over an item of property, usually to the exclusion of other parties. —From http://www.wikipedia.org/ The following illuminating description was obtained directly from Eugene Roshal, the designer of RAR. (The source code of the RAR decoder is available at [unrarsrc 06], but only on condition that it is not used to reverse-engineer the encoder.) RAR has two compression modes, general and special. The general mode employs an LZSS-based algorithm similar to ZIP Deflate (Section 6.25). The size of the sliding dictionary in RAR can be varied from 64 KB to 4 MB (with a 4 MB default value), and the minimum match length is 2. Literals, offsets, and match lengths are compressed further by a Huffman coder (recall that Deflate works similarly). 6.22 RAR and WinRAR 397 Starting at version 3.0, RAR also uses a special compression mode to improve the compression of text-like data. This mode is based on the PPMD algorithm (also known as PPMII) by Dmitry Shkarin. RAR contains a set of filters to preprocess audio (in wav or au formats), true color data, tables, and executables (32-bit x86 and Itanium, see note below). A data-analyzing module selects the appropriate filter and the best algorithm (general or PPMD), depending on the data to be compressed. (Note: The 80x86 family of processors was originally developed by Intel with a word length of 16 bits. Because of its tremendous success, its architecture was extended to 32-bit words and is currently referred to as IA-32 [Intel Architecture, 32-bit]. See [IA- 32 06] for more information. The Itanium is an IA-64 microprocessor developed jointly by Hewlett-Packard and Intel.) In addition to its use as Roshal ARchive, the acronym RAR has many other meanings, a few of which are listed here. (1) Remote Access Router (a network device used to connect remote sites via private lines or public carriers). (2) Royal Anglian Regiment, a unit of the British Army. (3) Royal Australian Regiment, a unit of the Australian Army. (4) Resource Adapter, a specific module of the Java EE platform. (5) The airport code of Rarotonga, Cook Islands. (6) Revise As Required (editors’ favorite). (7) Road Accident Rescue. See more at [acronyms 06]. Rarissimo, by [softexperience 06], is a file utility intended to automatically compress and decompress files in WinRAR. Rarissimo by itself is useless. It can be used only if WinRAR is already installed in the computer. The Rarissimo user specifies a number of folders for Rarissimo to watch, and Rarissimo compresses or decompresses any files that have been modified in these folders. It can also move the modified files to target folders. The user also specifies how often (in multiples of 10 sec) Rarissimo should check the folders. For each folder to be watched, the user has to specify RAR or UnRAR and a target folder. If RAR is specified, then Rarissimo employs WinRAR to compress each file that has been modified since the last check, and then moves that file to the target folder. If the target folder resides on another computer, this move amounts to an FTP transfer. If UnRAR has been specified, then Rarissimo employs WinRAR to decompress all the RAR-compressed files found in the folder and then moves them to the target folder. An important feature of Rarissimo is that it preserves NTFS alternate streams (see [ntfs 06]). This means that Rarissimo can handle Macintosh files that happen to reside on the PC; it can compress and decompress them while preserving their data and resource forks. 398 6. Dictionary Methods 6.23 The V.42bis Protocol The V.42bis protocol is a set of rules, or a standard, published by the ITU-T (page 248) for use in fast modems. It is based on the existing V.32bis protocol and is supposed to be used for fast transmission rates, up to 57.6K baud. Thomborson [Thomborson 92] in a detailed reference for this standard. The ITU-T standards are recommendations, but they are normally followed by all major modem manufacturers. The standard contains specifications for data compression and error correction, but only the former is discussed here. V.42bis specifies two modes: a transparent mode, where no compression is used, and a compressed mode using an LZW variant. The former is used for data streams that don’t compress well and may even cause expansion. A good example is an already compressed file. Such a file looks like random data; it does not have any repetitive patterns, and trying to compress it with LZW will fill up the dictionary with short, two-symbol, phrases. The compressed mode uses a growing dictionary, whose initial size is negotiated between the modems when they initially connect. V.42bis recommends a dictionary size of 2,048 entries. The minimum size is 512 entries. The first three entries, corresponding to pointers 0, 1, and 2, do not contain any phrases and serve as special codes. Code 0 (enter transparent mode—ETM) is sent when the encoder notices a low compression ratio, and it decides to start sending uncompressed data. (Unfortunately, V.42bis does not say how the encoder should test for low compression.) Code 1 is FLUSH, to flush data. Code 2 (STEPUP) is sent when the dictionary is almost full and the encoder decides to double its size. A dictionary is considered almost full when its size exceeds that of a special threshold (which is also negotiated by the modems). When the dictionary is already at its maximum size and it becomes full, V.42bis recommends a reuse procedure. The least-recently-used phrase is located and deleted, to make room for a new phrase. This is done by searching the dictionary from entry 256 for the first phrase that is not a prefix to any other phrase. Suppose that the phrase abcd is found, and there are no phrases of the form abcdx for any x. This means that abcd has not been used since it was created, and that it is the oldest such phrase. It therefore makes sense to delete it, since it reflects an old pattern in the input stream. This way, the dictionary always reflects recent patterns in the input. . . . there is an ever-increasing body of opinion which holds that The Ultra-Complete Maximegalon Dictionary is not worth the fleet of lorries it takes to cart its microstored edition around in. Strangely enough, the dictionary omits the word “floopily,” which simply means “in the manner of something which is floopy.” —Douglas Adams, Life, the Universe, and Everything (1982) 6.24 Various LZ Applications 399 6.24 Various LZ Applications ARC is a compression/archival/cataloging program developed by Robert A. Freed in the mid-1980s. It offers good compression and the ability to combine several files into one file, called an archive. Currently (early 2003) ARC is outdated and has been superseded by the newer PK applications. PKArc is an improved version of ARC. It was developed by Philip Katz, who has founded the PKWare company [PKWare 03], which still markets the PKzip, PKunzip, and PKlite software. The PK programs are faster and more general than ARC and also provide for more user control. Past editions of this book have more information on these applications. LHArc, from Haruyasu Yoshizaki, and LHA, by Haruhiko Okumura and Haruyasu Yoshizaki, use adaptive Huffman coding with features drawn from LZSS. The LZ window size may be up to 16K bytes. ARJ is another older compression tool by Robert K. Jung that uses the same algorithm, but increases the LZ window size to 26624 bytes. A similar program, called ICE (for the old MS-DOS operating system), seems to be a faked version of either LHarc or LHA. [Okumura 98] has some information about LHArc and LHA as well as a history of data compression in Japan. Two newer applications, popular with Microsoft Windows users, are RAR/WinRAR [rarlab 06] (Section 6.22) and ACE/WinAce [WinAce 03]. They use LZ with a large search buffer (up to 4 Mb) combined with Huffman codes. They are available for several platforms and offer many useful features. 6.25 Deflate: Zip and Gzip Deflate is a popular compression method that was originally used in the well-known Zip and Gzip software and has since been adopted by many applications, the most important of which are (1) the HTTP protocol ([RFC1945 96] and [RFC2616 99]), (2) the PPP compression control protocol ([RFC1962 96] and [RFC1979 96]), (3) the PNG (Portable Network Graphics, Section 6.27) and MNG (Multiple-Image Network Graphics) graphics file formats ([PNG 03] and [MNG 03]), and (4) Adobe’s PDF (Portable Document File, Section 11.13) [PDF 01]. Phillip W. Katz was born in 1962. He received a bachelor’s degree in computer science from the University of Wisconsin at Madison. Always interested in writing software, he started working in 1984 as a programmer for Allen-Bradley Co. developing programmable logic controllers for the industrial automation industry. He later worked for Graysoft, another software company, in Milwaukee, Wisconsin. At about that time he became interested in data compression and founded PKWare in 1987 to develop, implement, and market software products such as PKarc and PKzip. For a while, the company was very successful selling the programs as shareware. 400 6. Dictionary Methods Always a loner, Katz suffered from personal and legal problems, started drinking heavily, and died on April 14, 2000 from complications related to chronic alcoholism. He was 37 years old. After his death, PKWare was sold, in March 2001, to a group of investors. They changed its management and the focus of its business. PKWare currently targets the corporate market, and emphasizes compression combined with encryption. Their product line runs on a wide variety of platforms. Deflate was designed by Philip Katz as a part of the Zip file format and implemented in his PKZIP software [PKWare 03]. Both the ZIP format and the Deflate method are in the public domain, which allowed implementations such as Info-ZIP’s Zip and Unzip (essentially, PKZIP and PKUNZIP clones) to appear on a number of platforms. Deflate is described in [RFC1951 96]. The most notable implementation of Deflate is zlib, a portable and free compression library ([zlib 03] and [RFC1950 96]) by Jean-Loup Gailly and Mark Adler who designed and implemented it to be free of patents and licensing requirements. This library (the source code is available at [Deflate 03]) implements the ZLIB and GZIP file formats ([RFC1950 96] and [RFC1952 96]), which are at the core of most Deflate applications, including the popular Gzip software. Deflate is based on a variation of LZ77 combined with Huffman codes. We start with a simple overview based on [Feldspar 03] and follow with a full description based on [RFC1951 96]. The original LZ77 method (Section 6.3) tries to match the text in the look-ahead buffer to strings already in the search buffer. In the example search buffer look-ahead ...old..thea..then...there...|thenew......more the look-ahead buffer starts with the string the, which can be matched to one of three strings in the search buffer. The longest match has a length of 4. LZ77 writes tokens on the compressed stream, where each token is a triplet (offset, length, next symbol). The third component is needed in cases where no string has been matched (imagine having che instead of the in the look-ahead buffer) but it is part of every token, which reduces the performance of LZ77. The LZ77 algorithm variation used in Deflate eliminates the third component and writes a pair (offset, length) on the compressed stream. When no match is found, the unmatched character is written on the compressed stream instead of a token. Thus, the compressed stream consists of three types of entities: literals (unmatched characters), offsets (termed “distances” in the Deflate literature), and lengths. Deflate actually writes Huffman codes on the compressed stream for these entities, and it uses two code tables—one for literals and lengths and the other for distances. This makes sense because the literals are normally bytes and are therefore in the interval [0, 255], and the lengths are limited by Deflate to 258. The distances, however, can be large numbers because Deflate allows for a search buffer of up to 32Kbytes. When a pair (length, distance) is determined, the encoder searches the table of literal/length codes to find the code for the length. This code (we later use the term “edoc” for it) is then replaced by a Huffman code that’s written on the compressed stream. The encoder then searches the table of distance codes for the code of the 6.25 Deflate: Zip and Gzip 401 distance and writes that code (a special prefix code with a fixed, 5-bit prefix) on the compressed stream. The decoder knows when to expect a distance code, because it always follows a length code. The LZ77 variant used by Deflate defers the selection of a match in the following way. Suppose that the two buffers contain search buffer look-ahead ...old..sheneeds..then...there...|thenew......more input The longest match is 3. Before selecting this match, the encoder saves the t from the look-ahead buffer and starts a secondary match where it tries to match henew... with the search buffer. If it finds a longer match, it outputs t as a literal, followed by the longer match. There is also a 3-valued parameter that controls this secondary match attempt. In the “normal” mode of this parameter, if the primary match was long enough (longer than a preset parameter), the secondary match is reduced (it is up to the implementor to decide how to reduce it). In the “high-compression” mode, the encoder always performs a full secondary match, thereby improving compression but spending more time on selecting a match. In the “fast” mode, the secondary match is omitted. Deflate compresses an input data file in blocks, where each block is compressed separately. Blocks can have different lengths and the length of a block is determined by the encoder based on the sizes of the various prefix codes used (their lengths are limited to 15 bits) and by the memory available to the encoder (except that blocks in mode 1 are limited to 65,535 bytes of uncompressed data). The Deflate decoder must be able to decode blocks of any size. Deflate offers three modes of compression, and each block can be in any mode. The modes are as follows: 1. No compression. This mode makes sense for files or parts of files that are incompressible (i.e., random) or already compressed, or for cases where the compression software is asked to segment a file without compression. A typical case is a user who wants to move an 8 Gb file to another computer but has only a DVD “burner.” The user may want to segment the file into two 4 Gb segments without compression. Commercial compression software based on Deflate can use this mode of operation to segment the file. This mode uses no code tables. A block written on the compressed stream in this mode starts with a special header indicating mode 1, followed by the length LEN of the data, followed by LEN bytes of literal data. Notice that the maximum value of LEN is 65,535. 2. Compression with fixed code tables. Two code tables are built into the Deflate encoder and decoder and are always used. This speeds up both compression and decompression and has the added advantage that the code tables don’t have to be written on the compressed stream. The compression performance, however, may suffer if the data being compressed is statistically different from the data used to set up the code tables. Literals and match lengths are located in the first table and are replaced by a code (called “edoc”) that is, in turn, replaced by a prefix code that’s output to the compressed stream. Distances are located in the second table and are replaced by special prefix codes that are output to the compressed stream. A block written on the compressed stream in this mode starts with a special header indicating mode 2, followed by the compressed data in the form of prefix codes for the literals and lengths, and special prefix codes for the distances. The block ends with a single prefix code for end-of-block. 402 6. Dictionary Methods 3. Compression with individual code tables generated by the encoder for the particular data that’s being compressed. A sophisticated Deflate encoder may gather statistics about the data as it compresses blocks, and may be able to construct improved code tables as it proceeds from block to block. There are two code tables, for literals/lengths and for distances. They again have to be written on the compressed stream, and they are written in compressed format. A block written by the encoder on the compressed stream in this mode starts with a special header, followed by (1) a compressed Huffman code table and (2) the two code tables, each compressed by the Huffman codes that preceded them. This is followed by the compressed data in the form of prefix codes for the literals, lengths, and distances, and ends with a single code for end-of-block. 6.25.1 The Details Each block starts with a 3-bit header where the first bit is 1 for the last block in the file and 0 for all other blocks. The remaining two bits are 00, 01, or 10, indicating modes 1, 2, or 3, respectively. Notice that a block of compressed data does not always end on a byte boundary. The information in the block is sufficient for the decoder to read all the bits of the compressed block and recognize the end of the block. The 3-bit header of the next block immediately follows the current block and may therefore be located at any position in a byte on the compressed stream. The format of a block in mode 1 is as follows: 1. The 3-bit header 000 or 100. 2. The rest of the current byte is skipped, and the next four bytes contain LEN and the one’s complement of LEN (as unsigned 16-bit numbers), where LEN is the number of data bytes in the block. This is why the block size in this mode is limited to 65,535 bytes. 3. LEN data bytes. The format of a block in mode 2 is different: 1. The 3-bit header 001 or 101. 2. This is immediately followed by the fixed prefix codes for literals/lengths and the special prefix codes of the distances. 3. Code 256 (rather, its prefix code) designating the end of the block. Mode 2 uses two code tables: one for literals and lengths and the other for distances. The codes of the first table are not what is actually written on the compressed stream, so in order to remove ambiguity, the term “edoc” is used here to refer to them. Each edoc is converted to a prefix code that’s output to the compressed stream. The first table allocates edocs 0 through 255 to the literals, edoc 256 to end-of-block, and edocs 257–285 to lengths. The latter 29 edocs are not enough to represent the 256 match lengths of 3 through 258, so extra bits are appended to some of those edocs. Table 6.38 lists the 29 edocs, the extra bits, and the lengths represented by them. What is actually written on the compressed stream is prefix codes of the edocs (Table 6.39). Notice that edocs 286 and 287 are never created, so their prefix codes are never used. We show later that Table 6.39 can be represented by the sequence of code lengths 8, 8, . . . , 8    144 , 9, 9, . . . , 9    112 , 7, 7, . . . , 7    24 , 8, 8, . . . , 8    8 , (6.1) 6.25 Deflate: Zip and Gzip 403 Extra Extra Extra Code bits Lengths Code bits Lengths Code bits Lengths 257 0 3 267 1 15,16 277 4 67–82 258 0 4 268 1 17,18 278 4 83–98 259 0 5 269 2 19–22 279 4 99–114 260 0 6 270 2 23–26 280 4 115–130 261 0 7 271 2 27–30 281 5 131–162 262 0 8 272 2 31–34 282 5 163–194 263 0 9 273 3 35–42 283 5 195–226 264 0 10 274 3 43–50 284 5 227–257 265 1 11,12 275 3 51–58 285 0 258 266 1 13,14 276 3 59–66 Table 6.38: Literal/Length Edocs for Mode 2. edoc Bits Prefix codes 0–143 8 00110000–10111111 144–255 9 110010000–111111111 256–279 7 0000000–0010111 280–287 8 11000000–11000111 Table 6.39: Huffman Codes for Edocs in Mode 2. but any Deflate encoder and decoder include the entire table instead of just the sequence of code lengths. There are edocs for match lengths of up to 258, so the look-ahead buffer of a Deflate encoder can have a maximum size of 258, but can also be smaller. Examples. If a string of 10 symbols has been matched by the LZ77 algorithm, Deflate prepares a pair (length, distance) where the match length 10 becomes edoc 264, which is written as the 7-bit prefix code 0001000. A length of 12 becomes edoc 265 followed by the single bit 1. This is written as the 7-bit prefix code 0001010 followed by 1. A length of 20 is converted to edoc 269 followed by the two bits 01. This is written as the nine bits 0001101|01. A length of 256 becomes edoc 284 followed by the five bits 11110. This is written as 11000101|11110. A match length of 258 is indicated by edoc 285 whose 8-bit prefix code is 11000110. The end-of-block edoc of 256 is written as seven zero bits. The 30 distance codes are listed in Table 6.40. They are special prefix codes with fixed-size 5-bit prefixes that are followed by extra bits in order to represent distances in the interval [1, 32768]. The maximum size of the search buffer is therefore 32,768, but it can be smaller. The table shows that a distance of 6 is represented by 00100|1, a distance of 21 becomes the code 01000|101, and a distance of 8195 corresponds to code 11010|000000000010. 6.25.2 Format of Mode-3 Blocks In mode 3, the encoder generates two prefix code tables, one for the literals/lengths and 404 6. Dictionary Methods Extra Extra Extra Code bits Distance Code bits Distance Code bits Distance 0 0 1 10 4 33–48 20 9 1025–1536 1 0 2 11 4 49–64 21 9 1537–2048 2 0 3 12 5 65–96 22 10 2049–3072 3 0 4 13 5 97–128 23 10 3073–4096 4 1 5,6 14 6 129–192 24 11 4097–6144 5 1 7,8 15 6 193–256 25 11 6145–8192 6 2 9–12 16 7 257–384 26 12 8193–12288 7 2 13–16 17 7 385–512 27 12 12289–16384 8 3 17–24 18 8 513–768 28 13 16385–24576 9 3 25–32 19 8 769–1024 29 13 24577–32768 Table 6.40: Thirty Prefix Distance Codes in Mode 2. the other for the distances. It uses the tables to encode the data that constitutes the block. The encoder can generate the tables in any way. The idea is that a sophisticated Deflate encoder may collect statistics as it inputs the data and compresses blocks. The statistics are used to construct better code tables for later blocks. A naive encoder may use code tables similar to the ones of mode 2 or may even not generate mode 3 blocks at all. The code tables have to be written on the compressed stream, and they are written in a highly-compressed format. As a result, an important part of Deflate is the way it compresses the code tables and outputs them. The main steps are (1) Each table starts as a Huffman tree. (2) The tree is rearranged to bring it to a standard format where it can be represented by a sequence of code lengths. (3) The sequence is compressed by run-length encoding to a shorter sequence. (4) The Huffman algorithm is applied to the elements of the shorter sequence to assign them Huffman codes. This creates a Huffman tree that is again rearranged to bring it to the standard format. (5) This standard tree is represented by a sequence of code lengths which are written, after being permuted and possibly truncated, on the output. These steps are described in detail because of the originality of this unusual method. Recall that the Huffman code tree generated by the basic algorithm of Section 5.2 is not unique. The Deflate encoder applies this algorithm to generate a Huffman code tree, then rearranges the tree and reassigns the codes to bring the tree to a standard form where it can be expressed compactly by a sequence of code lengths. (The result is reminiscent of the canonical Huffman codes of Section 5.2.8.) The new tree satisfies the following two properties: 1. The shorter codes appear on the left, and the longer codes appear on the right of the Huffman code tree. 2. When several symbols have codes of the same length, the (lexicographically) smaller symbols are placed on the left. The first example employs a set of six symbols A–F with probabilities 0.11, 0.14, 0.12, 0.13, 0.24, and 0.26, respectively. Applying the Huffman algorithm results in a tree similar to the one shown in Figure 6.41a. The Huffman codes of the six symbols 6.25 Deflate: Zip and Gzip 405 are 000, 101, 001, 100, 01, and 11. The tree is then rearranged and the codes reassigned to comply with the two requirements above, resulting in the tree of Figure 6.41b. The new codes of the symbols are 100, 101, 110, 111, 00, and 01. The latter tree has the advantage that it can be fully expressed by the sequence 3, 3, 3, 3, 2, 2 of the lengths of the codes of the six symbols. The task of the encoder in mode 3 is therefore to generate this sequence, compress it, and write it on the compressed stream. A C D B E F E C F A B D (a) (b) 000 001 101 100 101 110 111 0 0 0 0 0 1 1 1 1 1 100 01 11 00 01 0 0 0 0 0 1 1 1 1 1 Figure 6.41: Two Huffman Trees. The code lengths are limited to at most four bits each. Thus, they are integers in the interval [0, 15], which implies that a code can be at most 15 bits long (this is one factor that affects the Deflate encoder’s choice of block lengths in mode 3). The sequence of code lengths representing a Huffman tree tends to have runs of identical values and can have several runs of the same value. For example, if we assign the probabilities 0.26, 0.11, 0.14, 0.12, 0.24, and 0.13 to the set of six symbols A–F, the Huffman algorithm produces 2-bit codes for A and E and 3-bit codes for the remaining four symbols. The sequence of these code lengths is 2, 3, 3, 3, 2, 3. The decoder reads a compressed sequence, decompresses it, and uses it to reproduce the standard Huffman code tree for the symbols. We first show how such a sequence is used by the decoder to generate a code table, then how it is compressed by the encoder. Given the sequence 3, 3, 3, 3, 2, 2, the Deflate decoder proceeds in three steps as follows: 1. Count the number of codes for each code length in the sequence. In our example, there are no codes of length 1, two codes of length 2, and four codes of length 3. 2. Assign a base value to each code length. There are no codes of length 1, so they are assigned a base value of 0 and don’t require any bits. The two codes of length 2 therefore start with the same base value 0. The codes of length 3 are assigned a base value of 4 (twice the number of codes of length 2). The C code shown here (after [RFC1951 96]) was written by Peter Deutsch. It assumes that step 1 leaves the number of codes for each code length n in bl_count[n]. code = 0; bl_count[0] = 0; for (bits = 1; bits <= MAX_BITS; bits++) { code = (code + bl_count[bits-1]) << 1; next code[bits] = code; } 406 6. Dictionary Methods 3. Use the base value of each length to assign consecutive numerical values to all the codes of that length. The two codes of length 2 start at 0 and are therefore 00 and 01. They are assigned to the fifth and sixth symbols E and F. The four codes of length 3 start at 4 and are therefore 100, 101, 110, and 111. They are assigned to the first four symbols A–D. The C code shown here (by Peter Deutsch) assumes that the code lengths are in tree[I].Len and it generates the codes in tree[I].Codes. for (n = 0; n <= max code; n++) { len = tree[n].Len; if (len != 0) { tree[n].Code = next_code[len]; next_code[len]++; } } In the next example, the sequence 3, 3, 3, 3, 3, 2, 4, 4 is given and is used to generate a table of eight prefix codes. Step 1 finds that there are no codes of length 1, one code of length 2, five codes of length 3, and two codes of length 4. The length-1 codes are assigned a base value of 0. There are zero such codes, so the next group is assigned the base value of twice 0. This group contains one code, so the next group (length-3 codes) is assigned base value 2 (twice the sum 0+1). This group contains five codes, so the last group is assigned base value of 14 (twice the sum 2 + 5). Step 3 simply generates the five 3-bit codes 010, 011, 100, 101, and 110 and assigns them to the first five symbols. It then generates the single 2-bit code 00 and assigns it to the sixth symbol. Finally, the two 4-bit codes 1110 and 1111 are generated and assigned to the last two (seventh and eighth) symbols. Given the sequence of code lengths of Equation (6.1), we apply this method to generate its standard Huffman code tree (listed in Table 6.39). Step 1 finds that there are no codes of lengths 1 through 6, that there are 24 codes of length 7, 152 codes of length 8, and 112 codes of length 9. The length-7 codes are assigned a base value of 0. There are 24 such codes, so the next group is assigned the base value of 2(0+24) = 48. This group contains 152 codes, so the last group (length-9 codes) is assigned base value 2(48 + 152) = 400. Step 3 simply generates the 24 7-bit codes 0 through 23, the 152 8-bit codes 48 through 199, and the 112 9-bit codes 400 through 511. The binary values of these codes are listed in Table 6.39. “Must you deflate romantic rhetoric? Besides, the Astabigans have plenty of visitors from other worlds who will be viewing her.” —Roger Zelazny, Doorways in the Sand It is now clear that a Huffman code table can be represented by a short sequence (termed SQ) of code lengths (herein called CLs). This sequence is special in that it tends to have runs of identical elements, so it can be highly compressed by run-length encoding. The Deflate encoder compresses this sequence in a three-step process where the first step employs run-length encoding; the second step computes Huffman codes for the run lengths and generates another sequence of code lengths (to be called CCLs) for those Huffman codes. The third step writes a permuted, possibly truncated sequence of the CCLs on the compressed stream. 6.25 Deflate: Zip and Gzip 407 Step 1. When a CL repeats more than three times, the encoder considers it a run. It appends the CL to a new sequence (termed SSQ), followed by the special flag 16 and by a 2-bit repetition factor that indicates 3–6 repetitions. A flag of 16 is therefore preceded by a CL and followed by a factor that indicates how many times to copy the CL. Thus, for example, if the sequence to be compressed contains six consecutive 7’s, it is compressed to 7, 16, 102 (the repetition factor 102 indicates five consecutive occurrences of the same code length). If the sequence contains 10 consecutive code lengths of 6, it will be compressed to 6, 16, 112, 16, 002 (the repetition factors 112 and 002 indicate six and three consecutive occurrences, respectively, of the same code length). Experience indicates that CLs of zero are very common and tend to have long runs. (Recall that the codes in question are codes of literals/lengths and distances. Any given data file to be compressed may be missing many literals, lengths, and distances.) This is why runs of zeros are assigned the two special flags 17 and 18. A flag of 17 is followed by a 3-bit repetition factor that indicates 3–10 repetitions of CL 0. Flag 18 is followed by a 7-bit repetition factor that indicates 11–138 repetitions of CL 0. Thus, six consecutive zeros in a sequence of CLs are compressed to 17, 112, and 12 consecutive zeros in an SQ are compressed to 18, 012. The sequence of CLs is compressed in this way to a shorter sequence (to be termed SSQ) of integers in the interval [0, 18]. An example may be the sequence of 28 CLs 4, 4, 4, 4, 4, 3, 3, 3, 6, 6, 6, 6, 6, 6, 6, 6, 6, 6, 0, 0, 0, 0, 0, 0, 2, 2, 2, 2 that’s compressed to the 16-number SSQ 4, 16, 012, 3, 3, 3, 6, 16, 112, 16, 002, 17, 112, 2, 16, 002, or, in decimal, 4, 16, 1, 3, 3, 3, 6, 16, 3, 16, 0, 17, 3, 2, 16, 0. Step 2. Prepare Huffman codes for the SSQ in order to compress it further. Our example SSQ contains the following numbers (with their frequencies in parentheses): 0(2), 1(1), 2(1), 3(5), 4(1), 6(1), 16(4), 17(1). Its initial and standard Huffman trees are shown in Figure 6.42a,b. The standard tree can be represented by the SSQ of eight lengths 4, 5, 5, 1, 5, 5, 2, and 4. These are the lengths of the Huffman codes assigned to the eight numbers 0, 1, 2, 3, 4, 6, 16, and 17, respectively. Step 3. This SSQ of eight lengths is now extended to 19 numbers by inserting zeros in the positions that correspond to unused CCLs. Position: 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 CCL: 4 5 5 1 5 0 5 0 0 0 0 0 0 0 0 0 2 4 0 Next, the 19 CCLs are permuted according to Position: 16 17 18 0 8 7 9 6 10 5 11 4 12 3 13 2 14 1 15 CCL: 2 4 0 4 0 0 0 5 0 0 0 5 0 1 0 5 0 5 0 The reason for the permutation is to end up with a sequence of 19 CCLs that’s likely to have trailing zeros. The SSQ of 19 CCLs minus its trailing zeros is written on the compressed stream, preceded by its actual length, which can be between 4 and 19. Each CCL is written as a 3-bit number. In our example, there is just one trailing zero, so the 18-number sequence 2, 4, 0, 4, 0, 0, 0, 5, 0, 0, 0, 5, 0, 1, 0, 5, 0, 5 is written on the compressed stream as the final, compressed code of one prefix-code table. In mode 3, 408 6. Dictionary Methods (a) (b) 0000 1100 00010 00011 1 0 00100 00101 11100 11101 11110 11111 0011 1101 01 10 0 0 1 2 1 2 3 3 4 6 4 6 16 16 17 17 Figure 6.42: Two Huffman Trees for Code Lengths. each block of compressed data requires two prefix-code tables, so two such sequences are written on the compressed stream. A reader finally reaching this point (sweating profusely with such deep concentration on so many details) may respond with the single word “insane.” This scheme of Phil Katz for compressing the two prefix-code tables per block is devilishly complex and hard to follow, but it works! The format of a block in mode 3 is as follows: 1. The 3-bit header 010 or 110. 2. A 5-bit parameter HLIT indicating the number of codes in the literal/length code table. This table has codes 0–256 for the literals, code 256 for end-of-block, and the 30 codes 257–286 for the lengths. Some of the 30 length codes may be missing, so this parameter indicates how many of the length codes actually exist in the table. 3. A 5-bit parameter HDIST indicating the size of the code table for distances. There are 30 codes in this table, but some may be missing. 4. A 4-bit parameter HCLEN indicating the number of CCLs (there may be between 4 and 19 CCLs). 5. A sequence of HCLEN + 4 CCLs, each a 3-bit number. 6. A sequence SQ of HLIT + 257 CLs for the literal/length code table. This SQ is compressed as explained earlier. 7. A sequence SQ of HDIST + 1 CLs for the distance code table. This SQ is compressed as explained earlier. 8. The compressed data, encoded with the two prefix-code tables. 9. The end-of-block code (the prefix code of edoc 256). Each CCL is written on the output as a 3-bit number, but the CCLs are Huffman codes of up to 19 symbols. When the Huffman algorithm is applied to a set of 19 symbols, the resulting codes may be up to 18 bits long. It is the responsibility of the encoder to ensure that each CCL is a 3-bit number and none exceeds 7. The formal definition [RFC1951 96] of Deflate does not specify how this restriction on the CCLs is to be achieved. 6.25.3 The Hash Table This short section discusses the problem of locating a match in the search buffer. The buffer is 32 Kb long, so a linear search is too slow. Searching linearly for a match to any string requires an examination of the entire search buffer. If Deflate is to be able to compress large data files in reasonable time, it should use a sophisticated search method. 6.25 Deflate: Zip and Gzip 409 The method proposed by the Deflate standard is based on a hash table. This method is strongly recommended by the standard, but is not required. An encoder using a different search method is still compliant and can call itself a Deflate encoder. Those unfamiliar with hash tables should consult any text on data structures. Instead of separate look-ahead and search buffers, the encoder should have one, 32 Kb buffer. The buffer is filled up with input data and initially all of it is a lookahead buffer. In the original LZ77 method, once symbols have been examined, they are moved into the search buffer. The Deflate encoder, in contrast, does not move the data in its buffer and instead moves a pointer (or a separator) from left to right, to indicate the point where the look-ahead buffer ends and the search buffer starts. Short, 3-symbol strings from the look-ahead buffer are hashed and added to the hash table. After hashing a string, the encoder examines the hash table for matches. Assuming that a symbol occupies n bits, a string of three symbols can have values in the interval [0, 23n-1]. If 23n-1 isn’t too large, the hash function can return values in this interval, which tends to minimize the number of collisions. Otherwise, the hash function can return values in a smaller interval, such as 32 Kb (the size of the Deflate buffer). We demonstrate the principles of Deflate hashing with the 17-symbol string abbaabbaabaabaaaa 12345678901234567 Initially, the entire 17-location buffer is the look-ahead buffer and the hash table is empty 0 1 2 3 4 5 6 7 8 0 0 0 0 0 0 0 0 ... We assume that the first triplet abb hashes to 7. The encoder outputs the raw symbol a, moves this symbol to the search buffer (by moving the separator between the two buffers to the right), and sets cell 7 of the hash table to 1. a|bbaabbaabaabaaaa 1 2345678901234567 0 1 2 3 4 5 6 7 8 0 0 0 0 0 0 0 1 ... The next three steps hash the strings bba, baa, and aab to, say, 1, 5, and 0. The encoder outputs the three raw symbols b, b, and a, moves the separator, and updates the hash table as follows: abba|abbaabaabaaaa 1234 5678901234567 0 1 2 3 4 5 6 7 8 4 2 0 0 0 3 0 1 ... Next, the triplet abb is hashed, and we already know that it hashes to 7. The encoder finds 1 in cell 7 of the hash table, so it looks for a string that starts with abb at position 1 of its buffer. It finds a match of size 6, so it outputs the pair (5 - 1, 6). The offset (4) is the difference between the start of the current string (5) and the start of the matching string (1). There are now two strings that start with abb, so cell 7 should point to both. It therefore becomes the start of a linked list (or chain) whose data items are 5 and 1. Notice that the 5 precedes the 1 in this chain, so that later searches of the chain will find the 5 first and will therefore tend to find matches with the smallest offset, because those have short Huffman codes. abbaa|bbaabaabaaaa 12345 678901234567 0 1 2 3 4 5 6 7 8 4 2 0 0 0 3 0 v ... 5 ›1 0 Six symbols have been matched at position 5, so the next position to consider is 6+5 = 11. While moving to position 11, the encoder hashes the five 3-symbol strings it 410 6. Dictionary Methods finds along the way (those that start at positions 6 through 10). They are bba, baa, aab, aba, and baa. They hash to 1, 5, 0, 3, and 5 (we arbitrarily assume that aba hashes to 3). Cell 3 of the hash table is set to 9, and cells 0, 1, and 5 become the starts of linked chains. abbaabbaab|aabaaaa 1234567890 1234567 0 1 2 3 4 5 6 7 8 v v 0 9 0 v 0 v ... . . . . . . 5 ›1 0 Continuing from position 11, string aab hashes to 0. Following the chain from cell 0, we find matches at positions 4 and 8. The latter match is longer and matches the 5-symbol string aabaa. The encoder outputs the pair (11 - 8, 5) and moves to position 11 + 5 = 16. While doing so, it also hashes the 3-symbol strings that start at positions 12, 13, 14, and 15. Each hash value is added to the hash table. (End of example.) It is clear that the chains can become very long. An example is an image file with large uniform areas where many 3-symbol strings will be identical, will hash to the same value, and will be added to the same cell in the hash table. Since a chain must be searched linearly, a long chain defeats the purpose of a hash table. This is why Deflate has a parameter that limits the size of a chain. If a chain exceeds this size, its oldest elements should be truncated. The Deflate standard does not specify how this should be done and leaves it to the discretion of the implementor. Limiting the size of a chain reduces the compression quality but can reduce the compression time significantly. In situations where compression time is unimportant, the user can specify long chains. Also, selecting the longest match may not always be the best strategy; the offset should also be taken into account. A 3-symbol match with a small offset may eventually use fewer bits (once the offset is replaced with a variable-length code) than a 4-symbol match with a large offset. 6.25.4 Conclusions Deflate is a general-purpose lossless compression algorithm that has proved valuable over the years as part of several popular compression programs. The method requires memory for the look-ahead and search buffers and for the two prefix-code tables. However, the memory size needed by the encoder and decoder is independent of the size of the data or the blocks. The implementation is not trivial, but is simpler than that of some modern methods such as JPEG 2000 or MPEG. Compression algorithms that are geared for specific types of data, such as audio or video, may perform better than Deflate on such data, but Deflate normally produces compression factors of 2.5 to 3 on text, slightly less for executable files, and somewhat more for images. Most important, even in the worst case, Deflate expands the data by only 5 bytes per 32 Kb block. Finally, free implementations that avoid patents are available. Notice that the original method, as designed by Phil Katz, has been patented (United States patent 5,051,745, September 24, 1991) and assigned to PKWARE. 6.26 LZMA and 7-Zip 411 6.26 LZMA and 7-Zip LZMA is the main (as well as the default) algorithm used in the popular 7z (or 7-Zip) compression software [7z 06]. Both 7z and LZMA are the creations of Igor Pavlov. The software runs on Windows and is free. Both LZMA and 7z were designed to provide high compression, fast decompression, and low memory requirements for decompression. The main feature of 7z is its open architecture. The software can currently compress data in one of six algorithms, and can in principle support any new compression methods. The current algorithms are the following: 1. LZMA. This is a sophisticated extension of LZ77 2. PPMD. A variant of Dmitry Shkarin’s PPMdH 3. BCJ. A converter for 32-bit x86 executables (see note) 4. BCJ2. A similar converter 5. BZip2. An implementation of the Burrows-Wheeler method (Section 11.1) 6. Deflate. An LZ77-based algorithm (Section 6.25) (Note: The 80x86 family of processors was originally developed by Intel with a word length of 16 bits. Because of its tremendous success, its architecture had been extended to 32-bit words and is currently referred to as IA-32 [Intel Architecture, 32-bit]. See [IA-32 06] for more information.) Other important features of 7z are the following: In addition to compressing and decompressing data in LZMA, the 7z software can compress and decompress data in ZIP, GZIP, and BZIP2, it can pack and unpack data in the TAR format, and it can decompress data originally compressed in RAR, CAB, ARJ, LZH, CHM, Z, CPIO, RPM, and DEB. When compressing data in the ZIP or GZIP formats, 7z provides compression ratios that are 2–10% better than those achieved by PKZip and WinZip. A data file compressed in 7z includes a decompressor; it is self-extracting. The 7z software is integrated with Windows Shell [Horstmann 06]. It constitutes a powerful file manager. It offers a powerful command line version. It has a plugin for FAR Manager. It includes localizations for 60 languages. It can encrypt the compressed data with the advanced encryption standard (AES- 256) algorithm, based on a 256-bit encryption key [FIPS197 01]. (The original AES algorithm, also known as Rijndael, was based on 128-bit keys.) The user inputs a text string as a passphrase, and 7z employs a key-derivation function to obtain the 256- bit key from the passphrase. This function is based on the SHA-256 hash algorithm [SHA256 02] and it computes the key by generating a very long string based on the passphrase and using it as the input to the hash function. The 256 bits output by the hash function become the encryption key. 412 6. Dictionary Methods To generate the string, 7z starts with the passphrase, encoded in the UTF-16 encoding scheme (two bytes per character, see [UTF16 06]). It then generates and concatenates 256K = 218 pairs (passphrase, integer) with 64-bit integers ranging from 0 to 218 - 1. If the passphrase is p symbols long, then each pair is 2p + 8 bytes long (the bytes are arranged in little endian), and the total length of the final string is 218(2p + 8) bytes; very long! The term Little Endian means that the low-order byte (the little end) of a number or a string is stored in memory at the lowest address (it comes first). For example, given the 4-byte number b3b2b1b0, if the least-significant byte b0 is stored at address A, then the most-significant byte b3 will be stored at address A + 3. LZMA (which stands for Lempel-Ziv-Markov chain-Algorithm) is the default compression algorithm of 7z. It is an LZ77 variant designed for high compression ratios and fast decompression. A free software development kit (SDK) with the LZMA source code, in C, C++, C#, and Java, is available at [7z 06] to anyone who wants to study, understand, or modify this algorithm. The main features of LZMA are the following: Compression speeds of about 1 Mb/s on a 2 GHz processor. Decompression speeds of about 10–20 Mb/s are typically obtained on a 2 GHz CPU The size of the decompressor can be as little as 2 Kb (if optimized for size of code), and it requires only 8–32Kb (plus the dictionary size) for its data. The dictionary size is variable and can be up to 4 Gb, but the current implementation limits it to 1 Gb. It supports multithreading and the Pentium 4’s hyperthreading. (Hyperthreading is a technology that allows resource sharing and partitioning while providing a multiprocessor environment in a unique CPU.) The current LZMA encoder can use one or two threads, and the LZMA decoder can use only one thread. The LZMA decoder uses only integer operations and can easily be implemented on any 32-bit processor (implementing it on a 16-bit CPU is more involved). The compression principle of LZMA is similar to that of Deflate (Section 6.25), but uses range encoding (Section 5.10.1) instead of Huffman coding. This complicates the encoder, but results in better compression (recall that range encoding is an integerbased version of arithmetic coding and can compress very close to the entropy of the data), while minimizing the number of renormalizations needed. Range encoding is implemented in binary such that shifts are used to divide integers, thereby avoiding the slow “divide” operation. Recall that LZ77 searches the search buffer for the longest string that matches the look-ahead buffer, then writes on the compressed stream a triplet (distance, length, next symbol) where “distance” is the distance from the string in the look-ahead buffer to its match in the search buffer. Thus, three types of data are written on the output, literals (the next symbol, often an ASCII code), lengths (relatively small numbers), and distances (which can be large numbers if the search buffer is large). LZMA also outputs these three types. If nothing matches the look-ahead buffer, a literal (a value in the interval [0, 255]) is output. If a match is found, then a pair (length, 6.26 LZMA and 7-Zip 413 distance) is output (after being encoded by the range coder). Because of the large size of the search buffer, a short, 4-entry, distance-history array is used that always contains the four most-recent distances that have been determined and output. If the distance of the current match equals one of the four distances in this array, then a pair (length, index) is output (after being encoded by the range coder), where “index” is a 2-bit index to the distance-history array. Locating matches in the search buffer is done by hashing two bytes, the current byte in the look-ahead buffer and the byte immediately to its right (but see the next paragraph for more details). The output of the hash function is an index to an array (the hash-array). The size of the array is selected as the power of 2 that’s closest to half the dictionary size, so the output of the hash function depends on the size of the dictionary. For example, if the dictionary size is 256 Mb (or 228), then the size of the array is 227 and the hash function has to compute and output 27-bit numbers. The large size of the array minimizes hash collisions. Experienced readers may have noticed the problem with the previous paragraph. If only two bytes are hashed, then the input to the hash function is only 16 bits, so there can be only 216 = 65,536 different inputs. The hash function can therefore compute only 65,536 distinct outputs (indexes to the hash-array) regardless of the size of the array. The full story is therefore more complex. The LZMA encoder can hash 2, 3, or 4 bytes and the number of items in the hash-array is selected by the encoder depending on the size of the dictionary. For example, a 1-Gb (= 230 bytes) dictionary results in a hash-array of size 512M = 229 items (where each item in the hash-array is a 32-bit integer). In order to take advantage of such a large hash-array, the encoder hashes four bytes. Four bytes constitute 32 bits, which provide the hash function with 232 distinct inputs. The hash function converts each input to a 29-bit index. (Thus, many inputs are converted to the same index.) Table 6.44 lists several user options and shows how the user can control the encoder by setting the Match Finder parameter to certain values. The value bt4, for example, specifies the binary tree mode with 2-3-4 bytes hashing. For simplicity, the remainder of this description talks about hashing two bytes (see Table 6.44 for various hashing options). The output of the hash function is used as an index to a hash-array and the user can choose one of two search methods, hash-chain (the fast method) or binary tree (the efficient method). In the fast method, the output of the hash function is an index to a hash-array of lists. Each array location is the start of a list of distances. Thus, if the two bytes hashed are XY and the result of the hash is index 123, then location 123 of the hash-array is a list of distances to pairs XY in the search buffer. These lists can be very long, so LZMA checks only the first 24 distances. These correspond to the 24 most-recent occurrences of XY (the number 24 is a user-controlled parameter). The best of the 24 matches is selected, encoded, and output. The distance of this match is then added to the start of the list, to become the first match found when array location 123 is checked again. This method is reminiscent of match searching in LZRW4 (Section 6.12). In the binary tree method, the output of the hash function is an index to a hasharray of binary search trees (Section 6.4). Initially, the hash-array is empty and there are no trees. As data is read and encoded, trees are generated and grow, and each 414 6. Dictionary Methods data byte becomes a node in one of the trees. A binary tree is created for every pair of consecutive bytes in the original data. Thus, if the data contains goodday, then trees are generated for the pairs go, oo, od, d, d, da, and ay. The total number of nodes in those trees is eight (the size of the data). Notice that the two occurrences of o end up as nodes in different trees. The first resides in the tree for oo, and the second ends up in the tree of od. The following example illustrates how this method employs the binary trees to find matches and how the trees are unpdated. The example assumes that the match-finder parameter (Table 6.44) is set to bt2, indicating 2-byte hashing. We assume that the data to be encoded is already fully stored in a long buffer (the dictionary). Five strings that start with the pair ab are located at various points as shown ...abm...abcd2...abcx...abcd1...aby... 1 2 3 5 7 1 4 0 7 8 We denote by p(n) the index (location) of string n in the dictionary. Thus, p(abm...) is 11 and p(abcd2) is 24. Each time a binary tree T is searched and a match is selected, T is rearranged and is updated according to the following two rules: 1. The tree must always remain a binary search tree. 2. If p(n1) < p(n2), then n2 cannot be in any subtree of n1. This implies that (a) the latest string (the one with the greatest index) is always the root of the tree and (b) indexes decrease as we slide down the tree. The result is a tree where recent strings are located near the root. We also assume that the pair ab of bytes is hashed to 62. Figure 6.43 illustrates how the binary tree for the pair ab is created, kept up to date, and searched. The following numbered items refer to the six parts of this figure. 1. When the LZMA encoder gets to location 11 and finds a, it hashes this byte and the b that follows, to obtain 62. It examines location 62 of the hash-array and finds it empty. It then generates a new binary tree (for the pair ab) with one node containing the pointer 11, and sets location 62 of the hash-array to point to this tree. There is no match, the byte a at 11 is output as a literal, and the encoder proceeds to hash the next pair bm. 2. The next pair ab is found by the LZMA encoder when it gets to location 24. Hashing produces the same 62, and location 62 of the hash-array is found to point to a binary tree whose root (which is so far its only node) contains 11. The encoder places 24 as the new root with 11 as its right subtree, because abcd2... is smaller than abm... (strings are compared lexicographically). The encoder matches abcd2... with abm... and the match length is 2. The encoder continues with cd2..., but before it does that, it hashes the pair bc and either appends it to the binary tree for bc (if such a tree exists) or generates a new tree. 3. The next pair ab is found at location 30. Hashing produces the same 62. The encoder places 30 (abcx...) as the new root with 24 (abcd2...) as its left subtree and 11 (abm..) as its right subtree. The better match is abcd2..., and the match length is 3. The next two pairs bc and cx are appended to their respective trees (if such trees exist), and the encoder continues with the pair x.. 6.26 LZMA and 7-Zip 415 4. The next occurrence of ab is found at location 57 and is hashed to 62. It becomes the root of the tree, with 30 (abcx...) as its right subtree. The best match is with abcd2... where the match length is 4. The encoder continues with the string 1... found at location 61. 5. In the last step of this example, the encoder finds a pair ab at location 78. This string (aby...) becomes the new root, with 57 as its left subtree. The match length is 2. 6. Now assume that location 78 contains the string abk.. instead of aby... Since abm.. is greater than abk.., it must be moved to the right subtree of abk.., resulting in a completely different tree. (1) (2) (3) (4) (5) (6) 62 11 abm.. 62 11 abcd2.. abm.. 24 62 24 abcd2.. abcx.. abm.. 11 30 62 abcd1.. abcx.. 57 62 aby.. 78 62 abk.. 78 24 abcd2.. abm.. 11 30 abcd1.. abcx.. 57 24 11 abm.. abm.. 11 30 57 24 30 Figure 6.43: Binary Search Trees in LZMA. Past versions of LZMA used a data structure called a Patricia trie (see page 357 for the definition of a trie), but this structure proved inefficient and has been eliminated. We end this description with some of the options that can be specified by the user when LZMA is invoked. -a{N}. The compression mode. 0, 1, and 2 specify the fast, normal, and max modes, with 2 as the default. (The latest version, currently in its beta stage, does not support the max mode.) -d{N}. The Dictionary size. N is an integer between 0 and 30, and the dictionary size is set to 2N. The default is N = 23 (an 8-Mb dictionary), and the current maximum is N = 30 (a 1-Gb dictionary). -mf{MF_ID}. The Match Finder. It specifies the method for finding matches and limits the number of bytes to be hashed. The memory requirements depend on this choice and on the dictionary size d. Table 6.44 lists the choices. The default value of MF_ID in the normal, max, and ultra modes is bt4 and in the fast and fastest modes it is hc4. We would like to thank Igor Pavlov for contributing important information and details to the material of this section. 416 6. Dictionary Methods MF ID Memory Description bt2 d Χ 9.5 + 1 Mb Binary Tree with 2 bytes hashing bt3 d Χ 11.5 + 4 Mb Binary Tree with 3 bytes hashing bt4 d Χ 11.5 + 4 Mb Binary Tree with 4 bytes hashing hc4 d Χ 7.5 + 4 Mb Hash Chain with 4 bytes hashing Table 6.44: LZMA Match Finder Options. Notes for Table 6.44: 1. bt4 uses three hash tables with 210 items for hashing two bytes, with 216 items for hashing three bytes, and with a variable size to hash four bytes. Only the latter table points to binary trees. The other tables point to positions in the input buffer. 2. bt3 uses two hash tables, one with 210 items for hashing two bytes and the other with a variable size to hash three bytes 3. bt2 uses only one hash table with 216 items. 4. bt2 and bt3 also can find almost all the matches, but bt4 is faster. 6.27 PNG The portable network graphics (PNG) file format has been developed in the mid-1990s by a group (the PNG development group [PNG 03]) headed by Thomas Boutell. The project was started in response to the legal issues surrounding the GIF file format (Section 6.21). The aim of this project was to develop a sophisticated graphics file format that will be flexible, will support many different types of images, will be easy to transmit over the Internet, and will be unencumbered by patents. The design was finalized in October 1996, and its main features are as follows: 1. It supports images with 1, 2, 4, 8, and 16 bitplanes. 2. Sophisticated color matching. 3. A transparency feature with very fine control provided by an alpha channel. 4. Lossless compression by means of Deflate combined with pixel prediction. 5. Extensibility: New types of meta-information can be added to an image file without creating incompatibility with existing applications. Currently, PNG is supported by many image viewers and web browsers on various platforms. This subsection is a general description of the PNG format, followed by the details of the compression method it uses. A PNG file consists of chunks that can be of various types and sizes. Some chunks contain information that’s essential for displaying the image, and decoders must be able to recognize and process them. Such chunks are referred to as “critical chunks.” Other chunks are ancillary. They may enhance the display of the image or may contain metadata such as the image title, author’s name, creation and modification dates and times, etc. (but notice that decoders may choose not to process such chunks). New, useful types of chunks can also be registered with the PNG development group. A chunk consists of the following parts: (1) size of the data field, (2) chunk name, (3) data field, and (4) a 32-bit cyclical redundancy code (CRC, Section 6.32). Each chunk has a 4-letter name of which (1) the first letter is uppercase for a critical chunk and lowercase for an ancillary chunk, (2) the second letter is uppercase for standard 6.27 PNG 417 chunks (those defined by or registered with the PNG group) and lowercase for a private chunk (an extension of PNG), (3) the third letter is always uppercase, and (4) the fourth letter is uppercase if the chunk is “unsafe to copy” and lowercase if it is “safe to copy.” Any PNG-aware application will process all the chunks it recognizes. It can safely ignore any ancillary chunk it doesn’t recognize, but if it finds a critical chunk it cannot recognize, it has to stop and issue an error message. If a chunk cannot be recognized but its name indicates that it is safe to copy, the application may safely read and rewrite it even if it has altered the image. However, if the application cannot recognize an “unsafe to copy” chunk, it must discard it. Such a chunk should not be rewritten on the new PNG file. Examples of “safe to copy” are chunks with text comments or those indicating the physical size of a pixel. Examples of “unsafe to copy” are chunks with gamma/color correction data or palette histograms. The four critical chunks defined by the PNG standard are IHDR (the image header), PLTE (the color palette), IDAT (the image data, as a compressed sequence of filtered samples), and IEND (the image trailer). The standard also defines several ancillary chunks that are deemed to be of general interest. Anyone with a new chunk that may also be of general interest may register it with the PNG development group and have it assigned a public name (a second letter in uppercase). The PNG file format uses a 32-bit CRC (Section 6.32) as defined by the ISO standard 3309 [ISO 84] or ITU-T V.42 [ITU-T 94]. The CRC polynomial is x32 + x26 + x23 + x22 + x16 + x12 + x11 + x10 + x8 + x7 + x5 + x4 + x2 + x + 1. The particular calculation proposed in the PNG standard employs a precalculated table that speeds up the computation significantly. A PNG file starts with an 8-byte signature that helps software to identify it as PNG. This is immediately followed by an IHDR chunk with the image dimensions, number of bitplanes, color type, and data filtering and interlacing. The remaining chunks must include a PLTE chunk if the color type is palette, and one or more adjacent IDAT chunks with the compressed pixels. The file must end with an IEND chunk. The PNG standard defines the order of the public chunks, whereas private chunks may have their own ordering constraints. An image in PNG format may have one of the following five color types: RGB with 8 or 16 bitplanes, palette with 1, 2, 4, or 8 bitplanes, grayscale with 1, 2, 4, 8, or 16 bitplanes, RGB with alpha channel (with 8 or 16 bitplanes), and grayscale with alpha channel (also with 8 or 16 bitplanes). An alpha channel implements the concept of transparent color. One color can be designated transparent, and pixels of that color are not drawn on the screen (or printed). Instead of seeing those pixels, a viewer sees the background behind the image. The alpha channel is a number in the interval [0, 2bp-1], where bp is the number of bitplanes. Assuming that the background color is B, a pixel in the transparent color P is painted in color (1 - .)B + .P. This is a mixture of .% background color and (1 - .)% pixel color. Perhaps the most intriguing feature of the PNG format is the way it handles interlacing. Interlacing makes it possible to display a rough version of the image on the screen, then improve it gradually until it reaches its final, high-resolution form. The special interlacing used in PNG is called Adam 7 after its developer, Adam M. Costello. 418 6. Dictionary Methods PNG divides the image into blocks of 8Χ8 pixels each, and displays each block in seven steps. In step 1, the entire block is filled up with copies of the top-left pixel (the one marked “1” in Figure 6.45a). In each subsequent step, the block’s resolution is doubled by modifying half its pixels according to the next number in Figure 6.45a. This process is easiest to understand with an example, such as the one shown in Figure 6.45b. (a) (b) 1 6 4 6 2 6 4 6 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 5 6 5 6 5 6 5 6 3 6 4 6 3 6 4 6 5 6 5 6 5 6 5 6 Figure 6.45: Interlacing in PNG. PNG compression is lossless and is performed in two steps. The first step, termed delta filtering (or just filtering), converts pixel values to numbers by a process similar to the prediction used in the lossless mode of JPEG (Section 7.10.5). The filtering step calculates a “predicted” value for each pixel and replaces the pixel with the difference between the pixel and its predicted value. The second step employs Deflate to encode the differences. Deflate is the topic of Section 6.25, so only filtering needs be described here. 6.27 PNG 419 Filtering does not compress the data. It only transforms the pixel data to a format where it is more compressible. Filtering is done separately on each image row, so an intelligent encoder can switch filters from one image row to the next (this is called adaptive filtering). PNG specifies five filtering methods (the first one is simply no filtering) and recommends a simple heuristic for choosing a filtering method for each image row. Each row starts with a byte indicating the filtering method used in it. Filtering is done on bytes, not on complete pixel values. This is easy to understand in cases where a pixel consists of three bytes, specifying the three color components of the pixel. Denoting the three bytes by a, b, and c, we can expect ai and ai+1 to be correlated (and also bi and bi+1, and ci and ci+1), but there is no correlation between ai and bi. Also, in a grayscale image with 16-bit pixels, it makes sense to compare the most-significant bytes of two adjacent pixels and then the least-significant bytes. Experiments suggest that filtering is ineffective for images with fewer than eight bitplanes, and also for palette images, so PNG recommends no filtering in such cases. The heuristic recommended by PNG for adaptive filtering is to apply all five filtering types to the row of pixels and select the type that produces the smallest sum of absolute values of outputs. (For the purposes of this test, the filtered bytes should be considered signed differences.) The five filtering types are described next. The first type (type 0) is no filtering. Filtering type 1 (sub) sets byte Bi,j in row i and column j to the difference Bi,j -Bi-t,j , where t is the interval between a byte and its correlated predecessor (the number of bytes in a pixel). Values of t for the various image types and bitplanes are listed in Table 6.46. If i-t is negative, then nothing is subtracted, which is equivalent to having a zero pixel on the left of the row. The subtraction is done modulo 256, and the bytes subtracted are considered unsigned. Image Bit Interval type planes t Grayscale 1, 2, 4, 8 1 Grayscale 16 2 Grayscale with alpha 8 2 Grayscale with alpha 16 4 Palette 1, 2, 4, 8 1 RGB 8 3 RGB 16 6 RGB with alpha 8 4 RGB with alpha 16 8 Figure 6.46: Interval Between Bytes. Filtering type 2 (up) sets byte Bi,j to the difference Bi,j -Bi,j-1. The subtraction is done as in type 1, but if j is the top image row, no subtraction is done. Filtering type 3 (average) sets byte Bi,j to the difference Bi,j-Bi-t,j+Bi,j-1χ2. The average of the left neighbor and the top neighbor is computed and subtracted from the byte. Any missing neighbor to the left or above is considered zero. Notice that the sum Bi-t,j + Bi,j-1 may be up to 9 bits long. To guarantee that the filtering is lossless 420 6. Dictionary Methods and can be reversed by the decoder, the sum must be computed exactly. The division by 2 is equivalent to a right shift and brings the 9-bit sum down to 8 bits. Following that, the 8-bit average is subtracted from the current byte Bi,j modulo 256 and unsigned. Example. Assume that the current byte Bi,j = 112, its left neighbor Bi-t,j = 182, and its top neighbor Bi,j-1 = 195. The average is (182 + 195) χ 2 = 188. Subtracting (112-188) mod 256 yields -76 mod 256 or 256-76 = 180. Thus, the encoder sets Bi,j to 180. The decoder inputs the value 180 for the current byte, computes the average in the same way as the encoder to end up with 188, and adds (180 + 188) mod 256 to obtain 112. Filtering type 4 (Paeth) sets byte Bi,j to Bi,j-PaethPredictBi-t,j,Bi,j-1,Bi-t,j-1. PaethPredict is a function that uses simple rules to select one of its three parameters, then returns that parameter. Those parameters are the left, top, and top-left neighbors. The selected neighbor is then subtracted from the current byte, modulo 256 unsigned. The PaethPredictor function is defined by the following pseudocode: function PaethPredictor (a, b, c) begin ; a=left, b=above, c=upper left p:=a+b-c ;initial estimate pa := abs(p-a) ; compute distances pb := abs(p-b) ; to a, b, c pc := abs(p-c) ; return nearest of a,b,c, ; breaking ties in order a,b,c. if pa<=pb AND pa<=pc then return a else if pb<=pc then return b else return c end PaethPredictor must perform its computations exactly, without overflow. The order in which PaethPredictor breaks ties is important and should not be altered. This order (that’s different from the one given in [Paeth 91]) is left neighbor, neighbor above, upper-left neighbor. PNG is a single-image format, but the PNG development group has also designed an animated companion format named MNG (multiple-image network format), which is a proper superset of PNG. We are indebted to Cosmin TrutΈa for reviewing and correcting this subsection. Does the world really need yet another graphics format? We believe so. GIF is no longer freely usable,. . . it would not be all that much easier to implement than a whole new file format. (PNG is designed to be simple to implement, with the exception of the compression engine, which would be needed in any case.) We feel that this is an excellent opportunity to design a new format that fixes some of the known limitations of GIF. From the PNG standard, RFC 2083, 1999 6.28 XML Compression: XMill 421 6.28 XML Compression: XMill XMill is a special-purpose, efficient software application for the compression of XML (Extensible Markup Language) documents. Its description in this short section is based on [Liefke and Suciu 99], but more details and a free implementation are available from [XMill 03]. First, a few words about XML. XML is a markup language for documents containing structured information. A markup language is a mechanism to identify structures in a document. A short story or a novel may have very little structure. It may be divided into chapters and may also include footnotes, an introduction, and an epilogue. A cooking recipe has more structure. It starts with the name of the dish and its class (such as salads, soups, etc.). This is followed by the two main structural items: the ingredients and the preparation. The recipe may then describe the proper way to serve the dish and may end with notes, reviews, and comments. A business card is similarly divided into a number of short items. The XML standard [XML 03] defines a way to add markup to documents, and has proven very popular and successful. An XML file contains data represented as text and also includes tags that identify the types of (or that assign meaning to) various data items. HTML is also a markup language, familiar to many, but it is restrictive in that it defines only certain tags and their meanings. The

tag, for example, is defined in HTML and has a certain meaning (a certain header), but anyone wanting to use a tag of, say, , has first to convince the WWW consortium to define it and assign it a meaning, then wait for the various web browsers to support it. XML, in contrast, does not specify a set of tags but provides a facility to define tags and structural relationships between them. The meaning of the tags (their semantics) is later defined by applications that process XML documents or by style sheets. Here is a simple example of a business card in XML. Most tags specify the start and end of a certain data item; they are wrappers. A tag such as that has a trailing “/” is considered empty. It specifies something to the application that reads and processes the XML file, but that something does not require any extra data. Melvin Schwartzkopf Chief person, Monster Inc. mschwa@monster.com (212)555-1414 In summary, an XML file consists of markup and content. There are six types of markup that can occur in an XML document: elements, entity references, comments, processing instructions, marked sections, and document-type declarations. The contents can be any digital data. The main aim of the developers of XMill was to design and implement a specialpurpose XML encoder that will compress an XML file better than a typical compressor will compress just the data of that file. Given a data file A, suppose that a typical 422 6. Dictionary Methods compressor, such as gzip, compresses it to X. Now add XML tags to A, converting it to B and increasing its size in the process. When B is compressed by XMill, the resulting file, Y , should be smaller than X. As an example, the developers had tested XMill on a 98-Mb data file taken from SwissProt, a data base for protein structure. The file was initially compressed by gzip down to 16 Mb. The original file was then converted to XML which increased its size to 165 Mb and was compressed by gzip (to 19 Mb) and by XMill (to 8.6 Mb, albeit after fine-tuning XMill for the specific data in that file). However, since XMill is a special-purpose encoder, it is not expected to perform well on arbitrary data files. The design of XMill is based on the following principles: 1. By itself, XMill is not a compressor. Rather, it is an extensible tool for specifying and applying existing compressors to XML data items. XMill analyzes the XML file, then invokes different compressors to compress different parts of the file. The main compressor used by XMill is gzip, but XMill includes other compressors and can also be linked by the user to any existing compressor. 2. Separate the structure from the raw data. The XML tags and attributes are separated by XMill from the data in the input file and are compressed separately. The data is the contents of XML elements and the values of attributes. 3. Group together items that are related. XMill uses the concept of a container. In the above example of business cards, all the URLs are grouped in one container, all the names are grouped in a second container, and so on. Also, all the XML tags and attributes are grouped in the structure container. The user can control the contents of the container by providing container expressions on the XMill command line. 4. Use semantic compressors. A container may include data items of a certain type, such as telephone numbers or airport codes. A sophisticated user may have an encoder that compresses such items efficiently. The user may therefore direct XMill to use certain encoders for certain containers, thereby achieving excellent compression for the entire XML input file. XMill comes with several built-in encoders, but any encoder available to the user may be linked to XMill and will be used by it to compress any specified container. An important part of XMill is a concise language, termed the container expressions, that’s used to group data items in containers and to specify the proper encoders for the various containers. XMill was designed to prepare XML files for storage or transmission. Sometimes, an XML file is used in connection with a query processor, where the file has to be searched often. XMill is not a good choice for such an application. Another limitation of XMill is that it performs well only on large files. Any XML file smaller than about 20 Kb will be poorly compressed by XMill because XMill adds overhead in the form of bookkeeping to the compressed file. Small XML files, however, are common in situations where messages in XML format are exchanged between users or between applications. I do not know it—it is without name—it is a word unsaid, It is not in any dictionary, utterance, symbol. —Walt Whitman, Leaves of Grass, (1900) 6.29 EXE Compressors 423 6.29 EXE Compressors The LZEXE program is freeware originally written in the late 1980s by Fabrice Bellard as a special-purpose utility to compress EXE files (PC executable files). The idea is that an EXE file compressed by LZEXE can be decompressed and executed with one command. The decompressor does not write the decompressed file on the disk but loads it in memory, relocates addresses, and executes it! The decompressor uses memory that’s eventually used by the program being decompressed, so it does not require any extra RAM. In addition, the decompressor is very small compared with decompressors in self-extracting archives. The algorithm is based on LZ. It uses a circular queue and a dictionary tree for finding string matches. The position and size of the match are encoded by an auxiliary algorithm based on the Huffman method. Uncompressed bytes are kept unchanged, since trying to compress them any further would have entailed a much more complex and larger decompressor. The decompressor is located at the end of the compressed EXE file and is 330 bytes long (in version 0.91). The main steps of the decoder are as follows: 1. Check the CRC (Section 6.32) to ensure data reliability. 2. Locate itself in high RAM; then move the compressed code in order to leave sufficient room for the EXE file. 3. Decompress the code, check that it is correct, and adjust the segments if bigger than 64K. 4. Decompress the relocation table and update the relocatable addresses of the EXE file. 5. Run the program, updating the CS, IP, SS, and SP registers. The idea of EXE compressors, introduced by LZEXE, was attractive to both users and software developers, so a few more have been developed: 1. PKlite, from PKWare, is a similar EXE compressor that can also compress .COM files. 2. DIET, by Teddy Matsumoto, is a more general EXE compressor that can compress data files. DIET can act as a monitor, permanently residing in RAM, watching for applications trying to read files from the disk. When an application tries to read a DIET-compressed data file, DIET senses it and does the reading and decompressing in a process that’s transparent to the application. UPX is an EXE compressor started in 1996 by Markus Oberhumer and L΄aszl΄o Moln΄ar. The current version (as of June 2006) is 2.01. Here is a short quotation from [UPX 03]. UPX is a free, portable, extendable, high-performance executable packer for several different executable formats. It achieves an excellent compression ratio and offers very fast decompression. Your executables suffer no memory overhead or other drawbacks. 424 6. Dictionary Methods 6.30 Off-Line Dictionary-Based Compression The dictionary-based compression methods described in this chapter are distinct, but have one thing in common; they generate the dictionary as they go along, reading data and compressing it. The dictionary is not included in the compressed file and is generated by the decoder in lockstep with the encoder. Thus, such methods can be termed “online.” In contrast, the methods described in this section are also based on a dictionary, but can be considered “offline” because they include the dictionary in the compressed file. BPE. The first method is byte pair encoding (BPE). This is a simple compression method, due to [Gage 94], that often features only mediocre performance. It is described here because (1) it is an example of a multipass method (two-pass compression algorithms are common, but multipass methods are generally considered too slow) and (2) it eliminates only certain types of redundancy and should therefore be applied only to data files that feature this redundancy. (The next method, by [Larsson and Moffat 00], does not suffer from these drawbacks and is much more efficient.) BPE is both an example of an offline dictionary-based compression algorithm and a simple example (perhaps the simplest) of a grammar-based compression method. In addition, the BPE decoder is very small, which makes it ideal for applications where memory size is restricted. The BPE method is easy to describe. We assume that the data symbols are bytes and we use the term bigram for a pair of consecutive bytes. Each pass locates the mostcommon bigram and replaces it with an unused byte value. Thus, the method performs best on files that have many unused byte values, and one aim of this discussion is to point out the types of data that feature this kind of redundancy. First, however, a small example. Given the character set A, B, C, D, X, and Y and the data file ABABCABCD (where X and Y are unused bytes), the first pass identifies the pair AB as the most-common bigram and replaces each of its three occurrences with the single byte X. The result is XXCXCD. The second pass identifies the pair XC as the most-common bigram and replaces each of its two occurrences with the single byte Y. The result is XYYD, where no bigram occurs more than once. Bigrams that occur just once can also be replaced, if more unused byte values are available. However, each replacement rule must be appended to the dictionary and thus ends up being included in the compressed file. As a result, the BPE encoder stops when no bigram occurs more than once. What types of data tend to have many unused byte values? The first type that comes to mind is text. Currently (in late 2008), most text files use the well-known ASCII codes to encode text. An ASCII code occupies a byte, but only seven bits constitue the actual code. The eighth bit can be used for a parity check, but is often simply set to zero. Thus, we can expect 128 byte values out of the 256 possible ones to be unused in a typical ASCII text file. A quick glance at an ASCII code table shows that codes 0 through 32 (as well as code 127) are control codes. They indicate commands such as backspace, carriage return, escape, delete, and blank space. It therefore makes sense to expect only a few of those to appear in any given text file. The validity of these arguments can be checked by a simple test. The following Mathematica code prints the unused byte values in a given text file. scn = ReadList["Hobbit.txt", Byte]; btc = Table[0, {256}]; 6.30 Off-Line Dictionary-Based Compression 425 Do[btc[[scn[[i]]]] = btc[[scn[[i]]]] + 1, {i, 1, Length[scn]}]; btc dis = Table[0, {256}]; j = 0; Do[If[btc[[i]] == 0, {j = j + 1, dis[[j]] = i - 1}], {i, 1, 256}]; Take[dis, j] It was executed on three chapters from the book Data Compression: The Complete Reference (4th edition) and on Tolkien’s The Hobbit. The following results were obtained: Dcomp3: 0–7, 9–11, 13–30, 126–255. Dcomp4.1: 0–11, 13–30, 126–255. Dcomp5: 0–7, 9–11, 13–30, 126–255. Hobbit: 0–7, 9–11, 13–30, 34–36, 42, 46, 60, 63, 87, 89–92, 95, 122–123, 125–255. The results of this experiment are clear. More than half the byte values are unused. The 128 values 128 through 255 are unused and the only ASCII control characters used are BS, FF, US, and Space. Today, more and more text files are encoded in Unicode, but even such files should have many unused byte values. (Notice that most Unicodes are 16 bits, but there are also 8-bit and 32-bit Unicodes.) A typical Unicode text file in an alphabetic language consists of letters, digits, punctuation marks, and accented letters, so the total number of codes is around 100–150, leaving many unused byte values. As an example, the 128 Unicodes for Greek and Coptic [Greek-Unicode 07] are 0370 through 03FF, and these do not use byte values 00, 01, and 04 through 6F. Naturally, text files in an ideograph-based language, such as Hangul or Cuneiforms, easily use all 256 byte values. Grayscale images constitute another example of data where many byte values may be unused. A typical image in grayscale may have millions of pixels, each in one of 256 shades of gray, but many shades may be unused. An experiment with the Mathematica code above indicates 41 unused byte values in the well-known lena image (raw, 128Χ128, 1-byte pixels), and 35 unused shades of gray (out of 256) in the familiar, raw format, baboon image of the same resolution. On the other hand, the same images in color (in raw format, with three bytes per pixels) use all the 256 byte values and it is easy to see why. A typical color image may consist of 6–7 million pixels (this is typical for today’s digital cameras) and may use only a few thousand colors. However, each color occupies three bytes, so even if a certain color, say (r, g, b)=(108, 56, 213), is unused, there is a good chance that some pixels have color components 108, 56, or 213. Thus, Gage’s method makes sense for compressing text and grayscale images. If it is applied to other types of data files, a large file should be segmented into small sections with unused byte values in each. At any given time, the method looks only at one pair of bytes, but this algorithm also indirectly takes advantage of longer repeating patterns. Thus, an input file of the form abcdeabcdfabcdg compresses quite efficiently. The most-common byte pairs are ab, bc, and cd. If the algorithm selects the first pair and replaces it with the unused byte x, the file becomes xcdexcdfxcdg. If xc is next selected and is replaced by y, the result is ydeydfydg. Now the pair yd is replaced by z, to produce zezfzg. The compression factor is 15/6 = 2.5, comparable to (or even exceeding) more efficient methods. 426 6. Dictionary Methods The remainder of this section describes a possible implementation of this method (reference [Gage 94] includes C source code). To compress a file, the entire file must be input into a buffer in memory (if the file is too large, it should be compressed in segments). As an example, we consider an alphabet of the eight symbols a through h (coded as 0 through 7) and the input file ababcabcd (with symbols e through h unused). The program constructs the following dictionary (also referred to as a pair-table or a phrase-table), where each zero in the bottom row indicates an unused character. 0 1 2 3 4 5 6 7 a b c d e f g h 1 1 1 1 0 0 0 0 The first step is to locate the most-common bigram. The simplest approach is to set up a table of size 256 Χ 256, initialize it to all zeros, use the two bytes of the next input bigram as row and column indexes to the table, and increment the particular entry pointed to by this bigram. The implementation described in [Gage 94] is based on a hash table that hashes each pair of bytes into a 12-bit number. That number is used as an index to an array of size 212 = 4,096 and the array location pointed to by the index is incremented. This works if the number of bigrams in the data file is less than 4,096 (notice that it can be up to 256 Χ 256 = 65,536). Once the most-common bigram is located (ab in our example), it is replaced by an unused character (h). The file in the buffer becomes hhchcd and the pair-table is updated to 0 1 2 3 4 5 6 7 a b c d e f g a 1 1 1 1 0 0 0 b The last entry indicates that byte-pair ab has been replaced by h (code 7). The next (and last) pass identifies pair hc as the most common bigram and replaces it with unused symbol g. The file in the buffer becomes hggd and the pair-table is updated to 0 1 2 3 4 5 6 7 a b c d e f h a 1 1 1 1 0 0 c b The 7th entry indicates that bigram hc has been replaced by g (code 6). The pair-table consists of two types of entries. One type (type 1) has a binary flag in the bottom row indicating used or unused symbols (1 or 0 flags). This type is easy to identify because the element in the top row is identical to its index (thus, the first element a had code 0 and is in column 0, while the last element, also a, has code 0 but is in column 7). The other type (type 2) indicates pair substitutions, and entries of this type should be written on the compressed file. Notice that the two types of entries may be mixed and do not have to be contiguous. In our example, the pair-table consists of six type-1 entries followed by two type-2 entries and is written on the compressed file as the six bytes -6, h, c, 1, a, b, where the first three bytes indicate six irrelevant type-1 entries (corresponding to codes 0 through 5) followed by one type-2 entry hc (corresponding to code 6). The next three bytes indicate one type-2 entry ab (which corresponds to code 7). The encoding rule for this table is therefore the following: Each contiguous segment of n type-1 entries followed by a type-2 entry is encoded as -n followed by the two bytes 6.30 Off-Line Dictionary-Based Compression 427 of the type-2 entry. Each segment of m consecutive type-2 entries is encoded as the byte m followed by the 2m bytes of the entries. The last feature of the encoder has to do with replacing pairs of bytes in the buffer. When a pair of bytes xy is replaced by a single byte p, it (the pair) becomes p-, where the - indicates an empty byte. There may be several ways to handle empty bytes. The straightforward way is to move bytes and compact the buffer each time a pair is replaced by a single byte. This is extremely slow because it may involve many thousands of byte movements for each replacement. A better approach is to have an auxiliary array of flags that indicate which byte positions in the buffer are empty. If the size of the buffer is n bytes, the size of the auxiliary array should be n bits, one-eighth (or 12.5%) the buffer size. If the buffer contains the 16 bytes thela-tfea-ure, then the auxiliary array should have the two bytes 00000010|00001000, where the two 1’s indicate empty bytes. When the compressed file is written from the buffer to the output, only those bytes that correspond to zero bits in the auxiliary array should be written. Another way to handle empty bytes is to organize the buffer as a linked list and establish another list (initially empty) of empty bytes. When a byte becomes empty, pointers are updated to take this byte out of the buffer and append it to the list of empty bytes. It is now clear that encoding may not be very efficient, but on the other hand the method never expands the data (except for the overhead from the pair-table). If no byte values are unused, the data is simply written on the compressed file as is, with the addition of the pair-table. This should be compared to other, more efficient methods where the “wrong” type of data may cause significant expansion. Encoding is slow, requiring multiple passes and a large buffer. Decoding, on the other hand, is fast. The decoder first reads and expands the pair-table, where only the type-2 entries are relevant. Bytes are then read from the compressed file. If a byte is literal (i.e., if it does not appear in the type-2 entries, such as byte d in our example), it is written to the final output as is. Otherwise, the byte is one of the type-2 entries and it represents a pair. The pair is constructed and is pushed into a stack. If the stack is not empty, the next byte is popped from the stack and handled as described above (which may cause another byte pair to be pushed into the stack). If the stack is empty, the next byte is read from the compressed file. Being so simple, the decoder is also very small, which is an advantage (as has been mentioned earlier). Re-Pair, an efficient offline dictionary algorithm The performance of BPE depends on the number of unused byte values in the original data. The next offline compression method (recursive pairing, or Re-Pair, by [Larsson and Moffat 00]) is much more sophisticated. It features high compression factors, a fast, small decoder, and it does not depend on the existence of unused byte values. It also employs sophisticated data structures that make it possible to select the most-common bigram in each pass without having to scan the entire input each time. No unused byte values are required. In each pass, the most-common bigram is identified and is replaced by a new symbol. We first illustrate this idea symbolically, using the well-known line from the television film Yabba-Dabba Do! (based on the 1960– 66 animated sitcom The Flintstones). The original data is shown in lowercase and the new symbols are in uppercase. 428 6. Dictionary Methods Pair String yabbadabba,yabbadabbadabbado A ›ba yabAdabA,yabAdabAdabAdo B ›ab yBAdBA,yBAdBAdBAdo C ›BA yCdC,yCdCdCdo D ›C yDdC,yDdDdDdo E ›Dd yEC,yEEEo F ›yE FC,FEEo The compressed file consists of the final string FC,FEEo and the six-entry dictionary (or phrase-table). Notice that the final string contains one bigram, but replacing it with a new symbol would require an extra dictionary entry and would therefore contribute nothing to the compression (and may even cause expansion). The main question is how to add new symbols. Assuming that the original data consists of bytes and that all 256 byte values are used, how can we add new symbols? We can start by placing each 8-bit byte in a pair of bytes (16 bits). Two-byte symbols can have values from 0 to 216 - 1 = 65,535, so there is room for the original 256 byte values and 65,280 new symbols. The bigram replacement algorithm is then executed and a phrase-table is constructed. Once this is done, the algorithm knows how many new symbols are needed, and the symbols (the 256 original ones and the new ones) are replaced with variable-length codes. One way to assign reasonable variable-length codes to the symbols is to count the number of occurrences of each symbol in the final string and in the phrase-table and use this information to construct a set of Huffman codes. The table of Huffman codes becomes overhead that must be included in the compressed file for the decoder’s use. An alternative is to sort the list of symbols according to their frequencies of occurrence and assign them one of the many variable-length codes for the integers, such as the Golomb, Rice, or Elias codes. The most-common symbols are assigned the shortest codes and the overhead in this case is the sorting information. This information—which is a permutation of the symbols, each represented by its variable-length code—must be included in the compressed file for the decoder’s use. In either case, the compressed file consists of three parts, the final string (where each symbol is represented by its variable-length code), the phrase-table (which must also be compressed), and the overhead. The Nakamura Murashima Method The Nakamura Murashima method ([Nakamura and Murashima 91] and [Nakamura and Murashima 96]) also employs new symbols, but encodes the phrase-table differently. The authors propose two variants, dubbed SCT and SED. The former is straightforward. Given the 4-symbol alphabet a, b, c, and d and the source data string 6.30 Off-Line Dictionary-Based Compression 429 acabbadcaddbaaddcaaddb, SCT encodes it as follows: Pair String acabbadcaddbaaddcaaddb A ›ad acabbAcAdbaAdcaAdb B ›Ad acabbAcBbaBcaBb C ›ca aCbbAcBbaBCBb D ›Bb aCbbAcDaBCD SCT encodes the phrase-table in a simple way and prepends it to the output. It generates a string that includes two symbols for each table entry. The names of the new symbols are not included in this string and are assumed to be A, B, C, and so on. Thus, our phrase-table is encoded as the 8-symbol string adAdcaBb and the complete encoder output is the string adAdcaBb|aCbbAcDaBCD, where the vertical bar is a special separator symbol. The decoder reconstructs the phrase-table easily. It reads pairs of symbols, assigns the first pair to the new symbol A, the second pair to new symbol B, and so on until it reaches the separator. The SED method is more complex. It constructs a different phrase-table and its output is a string of (original and new) symbols where flags distinguish between symbols and phrase-table entries. Each symbol in the output string is preceded by such a flag (a bit). A single symbol (either from the original alphabet or a new symbol) is preceded by a zero, while a phrase-table entry is preceded by a 1. Thus, the already-familiar input string acabbadcaddbaaddcaaddb becomes 0a10c0a0b0b10a0d0c110ί0d0b0a0.0.0. (where Greek letters stand for the new symbols). Here is why. The first input symbol a becomes 0a, but the following pair ca appears several times, so it becomes 10c0a, where the “1” indicates a new symbol, .. The next pair bb appears only once, so it becomes 0b0b, but the following pair ad repeats several times, so it becomes 10a0d, where the “1” indicates the next new symbol ί. The single c that follows becomes 0c. Notice that the encoder does not process the pair ca, because the c is followed by the pair ad, which has already been replaced by ί. Thus, the next substring addb becomes ίdb, and then .b (where the new symbol . is add), and finally the new symbol . = .b = ίdb = addb. This is encoded as the string 1(10ί0d)0b (without the paretheses). The rest of the encoding is easy to follow. There remains the question of how to write a hybrid string (with bits and symbols mixed up) such as 0a10c0a0b0b10a0d0c110ί0d0b0a0.0.0. on the output. The authors of this algorithm mention three different methods that employ adaptive arithmetic coding and complete binary trees, but no details are given. I would like to acknowledge the help provided by Hirofumi Nakamura in the preparation of this section. 430 6. Dictionary Methods 6.31 DCA, Compression with Antidictionaries A dictionary-based compression method is based on a dictionary; a collection of bits and pieces of data found in the input. If the encoder locates the next input string a in the dictionary, it compresses a by emitting a pointer to its location in the dictionary. An antidictionary method is based on an inverse kind of knowledge. Such a method maintains an antidictionary with strings that do not appear in the input. Using the antidictionary, the encoder can often predict the next data symbol a with certainty, so that a doesn’t have to be output. This is how such a method results in compression. The decoder can mimic the steps of the encoder and can therefore determine many data symbols and place them in the output. This section is based on [Crochemore et al. 00], who dubbed their approach DCA (data compression, antidictionaries). Reference [Davidson and Ilie 04] takes the basic idea further and describes algorithms for fast encoding and decoding. Antidictionary: This book would explain the disadvantages and hazards of using the word you looked up, and try to persuade you not to have anything to do with it. For example, under “solution,” it could say, “A term widely used by apish technologyfloggers to avoid the need to think of actual descriptive words for products and services.” Imagine what it could say under “country music.” http://www.halfbakery.com/idea/Anti-Dictionary#1096876321 An antidictionary method employs a binary alphabet where the basic data symbols are 0 and 1. We denote the input data (a bitstring) by w and assume that there exists an antidictionary AD with bitstrings that are not found in w (forbidden bitstrings). Notice that AD does not have to contain every forbidden bitstring. The encoder scans the input w bit by bit from left to right. We denote the part that has been scanned so far (the prefix of the input) by v. Assume that at a certain point, v is the bitstring 10110 and the next (still unknown) bit to be scanned is b. The encoder examines all the suffixes s of v which are 0, 10, 110, 0110, and 10110. If any of the strings s0 (i.e., a suffix s followed by a zero) is found in the antidictionary, then it is a forbidden string and the encoder knows that b must be 1. Similarly, if any of the strings s1 is in the AD, then b must be 0. In either case, b can be omitted from the output because the decoder can determine its value by searching the antidictionary in lockstep with the encoder. If neither s0 nor s1 is located in AD, then b is read by the encoder, is appended to v, and is also output because the decoder will not be able to determine it. Given the input string w = 10110101, it is easy to construct the set F(w) of substrings in w (we denote the empty string by .) F(w) = (., 0, 1, 01, 10, 11, 011, 101, 110, 0101, 0110, 1010, 1011, 1101, . . . , 10110101). An antidictionary AD for w is any set of bitstrings that are not in F(w). Such bitstrings are forbidden in w. For example, (00, 000, 001, 010, 100, 111, 0000, 0001) is such a set. Assuming that such an antidictionary exists, both encoder and decoder can easily be described and understood. Figure 6.47 is a pseudocode listing of the encoder. 6.31 DCA, Compression with Antidictionaries 431 ADC_Encoder(w,AD, .) 1. v ‹ .; . ‹ .; 2. for a ‹ first to last bit of w 3. if for every suffix s of v, neither s0 nor s1 is in AD 4. then . ‹ ..a; /* no compression */ 5. endif; 6. v ‹ v.a; 7. endfor 8. return .; Figure 6.47: A Simple DCA Encoder. As an example, we assume the input string w = 10110101 and the antidictionary AD = (00, 000, 111, 11011). Table 6.48 lists the eight encoding steps, the values of . and v at the end of each step, and the suffixes that have to be checked against AD at each step. Notice that the suffixes checked in each step are those of v of the previous step, because v is updated (in step 6) after the suffix check. Each suffix found in the AD (i.e., each forbidden string) is underlined, and it is obvious that each step with an underlined suffix leaves . unchanged and thus increases compression (the difference in length between . and v) by one bit. At the end of the loop, variable v equals the original input w, which suggests that there is really no need to have v as an independent variable. v is always a prefix of w and it is enough to maintain a pointer to w that will point to the current bit (the rightmost bit of v) at each step. a . v Suffix pairs s0, s1 to examine Initial . . 1. 1 1 1 .0, .1 2. 0 10 10 10, 11 3. 1 10 101 00, 01, 100, 101 4. 1 101 1011 10, 11, 010, 011, 1010, 1011 5. 0 101 10110 10, 11, 110, 111, 0110, 0111, 10110, 10111 6. 1 101 101101 00, 01, 100, 101, 1100, 1101, 01100, 01101, 101100, . . . 7. 0 101 1011010 10, 11, 010, 011, 1010, 1011, 11010, 11011, 011010, . . . 8. 1 101 10110101 00, 01, 100, 101, 0100, 0101, 10100, 10101, 110100, . . . Table 6.48: DCA Encoding of w = 10110101. The compressed stream that is output by the encoder must consist of (1) the AD, (2) the length n of w, and (3) bitstring .. Notice that n is needed by the decoder because different input strings w may produce the same . when processed by the encoder of Figure 6.47 (this is illustrated by the decoding listing of Table 6.50). The antidictionary must also be included in the encoder’s output, because it is static and has to be used by the decoder. (Notice that the entire input string w must be input in order to construct the antidictionary, which implies that DCA is a two-pass method.) Naturally, including 432 6. Dictionary Methods AD in the compressed stream reduces the effectiveness of this method, and suggests that an algorithm to construct an effective AD should be a crucial part of any practical DCA implementation. Alternatively, a dynamic approach is also possible, where the AD starts empty and is populated by forbidden strings as the encoder (and in lockstep, also the decoder) scans more and more input bits. At the very least, the AD should be compressed before it is written on the output. Notice that an AD is specific to the input data, which is why a general AD, based on training documents, cannot be used. The DCA decoder is simple. It receives an AD, the length n of the original data w, and the compressed string .. The decoder loops n times. In each iteration, it has a prefix v of w from previous iterations. It examines all the suffixes s of v. If a bitstring s0 is in the AD, then the next reconstructed bit must be a 1. Similarly, if s1 is found in the antidictionary, then the next bit must be 0. The next bit is then appended to v. (The notation |v| denotes the length of string v, while b denotes the complement of bit b.) If neither s0 nor s1 is in the antidictionary, the next bit of . is appended to v. In either case, v grows by one bit in the iteration. Figure 6.49 is a pseudocode listing of this simple decoder. ADC_Decoder(AD, n, ., v) 1. v ‹ .; 2. while |v| < n 3. if for a suffix s of v and a bit b, string s.b is in AD 4. then v ‹ v.―b; 5. else v ‹ v.(next bit of .); 6. endif; 7. endwhile; 8. return v; Figure 6.49: A Simple DCA Decoder. Table 6.50 lists the steps performed by this decoder when it is given the same antidictionary and is asked to perform eight iterations with . = 101 (when a bit of . is used, it is underlined). Notice that the last (least-significant) bit of . is used in iteration 4, which is why the decoder has to be given the length of the final, reconstructed bitstring. Here are some observations regarding antidictionary design. If a substring s of the input w is forbidden, but a proper substring u of s is also forbidden, then u but not s should be included in the antidictionary. In general, only minimal forbidden strings (strings where no substring is also forbidden) should appear in an antidictionary. In our examples above, the antidictionary contains 000 but this string is never used because it is not minimal. Also, most matches between suffixes s0, s1 and antidictionary strings occur for short suffixes s. Thus, it is more important to include short forbidden strings in an AD, which keeps it short. In practice, the length of an input string may be many thousands of bits, whereas the length of the antidictionary may be a few hundred bits. Still, it is important to construct an antidictionary by an efficient algorithm and also to look for ways to compress the AD. 6.31 DCA, Compression with Antidictionaries 433 v |v| . Suffix pairs s0, s1 to examine 1. . ›1 0 101 none 2. 1 › 10 1 101 10, 11 3. 10 › 101 2 00, 01, 100, 101 4. 101 › 1011 3 101 10, 11, 010, 011, 1010, 1011 5. 1011 › 10110 4 10, 11, 110, 111, 0110, 0111, 10110, 10111 6. 10110 › 101101 5 00, 01, 100, 101, 1100, 1101, 01100, 01101, . . . 7. 101101 › 1011010 6 10, 11, 010, 011, 1010, 1011, 11010, 11011, . . . 8. 1011010 › 10110101 7 00, 01, 100, 101, 0100, 0101, 10100, 10101, . . . Table 6.50: DCA Decoding of . = 101 and n = 8. The main reference of DCA is [Crochemore et al. 00] and it discusses ways to prune the antidictionary and to compress it. The developers observe the following. Given an antidictionary AD with only minimal forbidden strings, if we remove a string v from AD, the remaining strings constitute an antidictionary for v. Thus, an antidictionary can be compressed by removing each of its strings in turn and then using the remaining antidictionary to compress the string; a process that can be termed self compression. The developers of this method also describe an algorithm to construct an AD. Unfortunately, this algorithm is complex, it is based on finite-state automata, and its construction is beyond the scope of this book. Instead, we show an example (after [Davidson and Ilie 04]) illustrating how the automaton that corresponds to an antidictionary for a given bitstring w can be used to create two state diagrams (also referred to as transducers) that simplify the encoding and decoding of w. Given the input string w = 11010001, the AD-creation algorithm (not described here) produces the automaton of Figure 6.51a. Each state in such an automaton has two outgoing transitions, for 0 and 1. If one of the transitions of state A leads to a non-final state (a square), then A is referred to as a predict state and is shown in gray in the figure. Once this automaton has been obtained, it is used to create the antidictionary AD = (0000, 111, 011, 0101, 1100) in the following way. We start at state 0 and work our way to one of the non-final states (shown as squares), collecting bits off the edges on our way and concatenating them to form a bitstring. When we get to a non-final state, the bitstring is added to AD as the next forbidden string. Once we have visited all the non-final states, the AD is complete. With the AD in hand, the same automaton is used to construct a state diagram (a transducer) for encoding w = 11010001 (Figure 6.51b). This transducer has the following property: For each outgoing transition from a non-predict state (a white circle), the output equals the input. Thus, in our example, we enter state 0 with an input of 1 (because w starts with 1), so we have to leave it with the same output, which takes us to state 2. The input bit at this point is the second 1 of w, so we leave state 2 with an identical output and find ourselves in state 5. The next input bit is 0, so the output is . and the transducer sends us to state 9. The entire encoding sequence can be summarized by 0 1/1 › 2 1/1 › 5 0/. › 9 1/. › 4 0/. › 7 0/. › 3 0/0 › 6 1/. › 4 434 6. Dictionary Methods                                                          Figure 6.51: Automaton and Transducer for Fast Encoding and Decoding. and the encoding produces (in states 0, 2, and 3) the output string 110. Notice that state 5 has only one output transition (corresponding to an input of 0). The reason is that we can arrive at this state only after scanning the substring 11 from the input, and the AD tells us that this substring cannot be followed by a 0. The same transducer is used for decoding. The only change needed is to swap the input and output bits on each transition edge. Thus, for example, the transition from state 5 to state 9 will now be labeled ./0. Decoding string 111 is done in eight steps (notice that the decoder requires the length of w) as follows 0 1/1 › 2 1/1 › 5 ./0 › 9 ./1 › 4 ./0 › 7 ./0 › 3 0/0 › 6 ./1 › 4. It seems that the idea of using antidictionaries for compression has caught the attention of many researchers. Here are just two interesting contributions to this area. Reference [Crochemore and Navarro 02] introduces the notion of almost antifactors, which are strings that rarely appear in the text. More precisely, almost antifactors are strings that improve compression if they are considered forbidden and are added to the antidictionary. In [Ota and Morita 07], the authors present an algorithm (linear in space and time) to construct an antidictionary from a suffix tree. 6.32 CRC The idea of a parity bit is simple, old, and familiar to most computer practitioners. A parity bit is the simplest type of error detecting code. It adds reliability to a group of bits by making it possible for hardware to detect certain errors that may occur when the group is stored in memory, is written on a disk, or is transmitted over communication lines between computers. A single parity bit does not make the group completely reliable. There are certain errors that cannot be detected with a parity bit, but experience shows that even a single parity bit can make data transmission reliable in most practical cases. The parity bit is computed from a group of n - 1 bits, then added to the group, making it n bits long. A common example is a 7-bit ASCII code that becomes 8 bits long after a parity bit is added. The parity bit p is computed by counting the number of 1’s in the original group, and setting p to complete that number to either odd or even. The former is called odd parity, and the latter is called even parity. 6.32 CRC 435 Examples: Given the group of 7 bits 1010111, the number of 1’s is five, which is odd. Assuming odd parity, the value of p should be 0, leaving the total number of 1’s odd. Similarly, the group 1010101 has four 1’s, so its odd parity bit should also be a 1, bringing the total number of 1’s to five. Imagine a block of data where the most significant bit (MSB) of each byte is an odd parity bit, and the bytes are written vertically (Table 6.52a). 1 01101001 0 00001011 0 11110010 0 01101110 1 11101101 1 01001110 0 11101001 1 11010111 (a) 1 01101001 0 00001011 0 11010010 0 01101110 1 11101101 1 01001110 0 11101001 1 11010111 (b) 1 01101001 0 00001011 0 11010110 0 01101110 1 11101101 1 01001110 0 11101001 1 11010111 (c) 1 01101001 0 00001011 0 11010110 0 01101110 1 11101101 1 01001110 0 11101001 1 11010111 0 00011100 (d) Table 6.52: Horizontal and Vertical Parities. When this block is read from a disk or is received by a computer, it may contain transmission errors, errors that have been caused by imperfect hardware or by electrical interference during transmission. We can think of the parity bits as horizontal reliability. When the block is read, the hardware can check every byte, verifying the parity. This is done by simply counting the number of 1’s in the byte. If this number is odd, the hardware assumes that the byte is good. This assumption is not always correct, since two bits may get corrupted during transmission (Table 6.52c). A single parity bit is therefore useful (Table 6.52b) but does not provide full error detection capability. A simple way to increase the reliability of a block of data is to compute vertical parities. The block is considered to be eight vertical columns, and an odd parity bit is computed for each column (Table 6.52d). If two bits in a byte go bad, the horizontal parity will not catch it, but two of the vertical ones will. Even the vertical bits do not provide complete error detection capability, but they are a simple way to significantly improve data reliability. A CRC is a glorified vertical parity. CRC stands for Cyclical Redundancy Check (or Cyclical Redundancy Code) and is a rule that shows how to compute the vertical check bits (they are now called check bits, not just simple parity bits) from all the bits of the data. Here is how CRC-32 (one of the many standards developed by the CCITT) is computed. The block of data is written as one long binary number. In our example this will be the 64-bit number 101101001|000001011|011110010|001101110|111101101|101001110|011101001|111010111. 436 6. Dictionary Methods The individual bits are considered the coefficients of a polynomial (see below for definition). In our example, this will be the degree-63 polynomial P(x) = 1 Χ x63 + 0 Χ x62 + 1 Χ x61 + 1 Χ x60 + · · · + 1 Χ x2 + 1 Χ x1 + 1 Χ x0 = x63 + x61 + x60 + · · · + x2 + x + 1. This polynomial is then divided by the standard CRC-32 generating polynomial CRC32(x) = x32+x26+x23+x22+x16+x12+x11+x10+x8+x7+x5+x4+x2+x1+1. When an integer M is divided by an integer N, the result is a quotient Q (which we will ignore) and a remainder R, which is in the interval [0,N - 1]. Similarly, when a polynomial P(x) is divided by a degree-32 polynomial, the result is two polynomials, a quotient and a remainder. The remainder is a degree-31 polynomial, which means that it has 32 coefficients, each a single bit. Those 32 bits are the CRC-32 code, which is appended to the block of data as four bytes. As an example, the CRC-32 of a recent version of the file with the text of this chapter is 586DE4FE16. Selecting a generating polynomial is more an art than science. Page 196 of [Tanenbaum 02] is one of a few places where the interested reader can find a clear discussion of this topic. The CRC is sometimes called the “fingerprint” of the file. Of course, since it is a 32-bit number, there are only 232 different CRCs. This number equals approximately 4.3 billion, so, in theory, there may be different files with the same CRC, but in practice this is rare. The CRC is useful as an error detecting-code because it has the following properties: 1. Every bit in the data block is used to compute the CRC. This means that changing even one bit may produce a different CRC. 2. Even small changes in the data normally produce very different CRCs. Experience with CRC-32 shows that it is very rare that introducing errors in the data does not modify the CRC. 3. Any histogram of CRC-32 values for different data blocks is flat (or very close to flat). For a given data block, the probability of any of the 232 possible CRCs being produced is practically the same. Other common generating polynomials are CRC12(x) = x12 + x3 + x + 1 and CRC16(x) = x16 + x15 + x2 + 1. They generate the common CRC-12 and CRC-16 codes, which are 12 and 16 bits long, respectively. Definition: A polynomial of degree n in x is the function Pn(x) = n i=0 aixi = a0 + a1x + a2x2 + · · · + anxn, where ai are the n + 1 coefficients (in our case, real numbers). Two simple, readable references on CRC are [Ramabadran and Gaitonde 88] and [Williams 93]. 6.33 Summary 437 6.33 Summary The dictionary-based methods presented here are different but are based on the same principle. They read the input stream symbol by symbol and add phrases to the dictionary. The phrases are symbols or strings of symbols from the input. The main difference between the methods is in deciding what phrases to add to the dictionary. When a string in the input stream matches a dictionary phrase, the encoder outputs the position of the match in the dictionary. If that position requires fewer bits than the matched string, compression results. In general, dictionary-based methods, when carefully implemented, give better compression than statistical methods. This is why many popular compression programs are dictionary based or employ a dictionary as one of several compression steps. 6.34 Data Compression Patents It is generally agreed that an invention or a process is patentable but a mathematical concept, calculation, or proof is not. An algorithm seems to be an abstract mathematical concept that should not be patentable. However, once the algorithm is implemented in software (or in firmware) it may not be possible to separate the algorithm from its implementation. Once the implementation is used in a new product (i.e., an invention), that product—including the implementation (software or firmware) and the algorithm behind it—may be patentable. [Zalta 88] is a general discussion of algorithm patentability. Several common data compression algorithms, most notably LZW, have been patented; and the LZW patent is discussed here in some detail. The Sperry Corporation was granted a patent (4,558,302) on LZW in December 1985 (even though the inventor, Terry Welch, left Sperry prior to that date). When Unisys acquired Sperry in 1986 it became the owner of this patent and required users to obtain (and pay for) a license to use it. When CompuServe designed the GIF format in 1987 it decided to use LZW as the compression method for GIF files. It seems that CompuServe was not aware at that point that LZW was patented, nor was Unisys aware that the GIF format uses LZW. After 1987 many software developers became attracted to GIF and published programs to create and display images in this format. GIF became widely accepted and was commonly used on the World-Wide Web, where it was one of the prime image formats for Web pages and browsers. It was not until GIF had become a world-wide de facto standard that Unisys contacted CompuServe for a license. Naturally, CompuServe and other LZW users tried to challenge the patent. They applied to the United States Patent Office for a reexamination of the LZW patent, with the (perhaps surprising) result that on January 4, 1994, the patent was reconfirmed and CompuServe had to obtain a license (for an undisclosed sum) from Unisys later that year. Other important licensees of LZW (see [Rodriguez 95]) were Aldus (in 1991, for the TIFF graphics file format), Adobe (in 1990, for PostScript level II), and America Online and Prodigy (in 1995). The Unisys LZW patent had significant implications for the World-Wide Web, where use of GIF format images was currently widespread. Similarly, the Unix compress 438 6. Dictionary Methods utility uses LZW and therefore required a license. In the United States, the patent expired on 20 June 2003 (20 years from the date of first filing). In Europe (patent EP0129439) expired on 18 June 2004. In Japan, patents 2,123,602 and 2,610,084 expired on 20 June 2004, and in Canada, patent CA1223965 expired on 7 July 2004. Unisys graciously exempted old software products (those written or modified before January 1, 1995) from a patent license. Also exempt was any noncommercial and nonprofit software, old and new. Commercial software (even shareware) or firmware created after December 31, 1994, had to be licensed if it supported the GIF format or implemented LZW. A similar policy was enforced with regard to TIFF, where the cutoff date is July 1, 1995. Notice that computer users could legally keep and transfer GIF and any other files compressed with LZW; only the compression/decompression software required a license. For more information on the Unisys LZW patent and license see [unisys 03]. An alternative to GIF is the Portable Network Graphics, PNG (pronounced “ping,” Section 6.27) graphics file format [Crocker 95], which was developed expressly to replace GIF, and avoid patent claims. PNG is simple, portable, with source code freely available, and is unencumbered by patent licenses. It has potential and promise in replacing GIF. However, any GIF-to-PNG conversion software required a Unisys license. The GNU gzip compression software (Section 6.14) should also be mentioned here as a popular substitute for compress, since it is free from patent claims, is faster, and provides superior compression. The LZW U.S. patent number is 4,558,302, issued on Dec. 10, 1985. Here is the abstract filed as part of it (the entire filing constitutes 50 pages). A data compressor compresses an input stream of data character signals by storing in a string table strings of data character signals encountered in the input stream. The compressor searches the input stream to determine the longest match to a stored string. Each stored string comprises a prefix string and an extension character where the extension character is the last character in the string and the prefix string comprises all but the extension character. Each string has a code signal associated therewith and a string is stored in the string table by, at least implicitly, storing the code signal for the string, the code signal for the string prefix and the extension character. When the longest match between the input data character stream and the stored strings is determined, the code signal for the longest match is transmitted as the compressed code signal for the encountered string of characters and an extension string is stored in the string table. The prefix of the extended string is the longest match and the extension character of the extended string is the next input data character signal following the longest match. Searching through the string table and entering extended strings therein is effected by a limited search hashing procedure. Decompression is effected by a decompressor that receives the compressed code signals and generates a string table similar to that constructed by the compressor to effect lookup of received code signals so as to recover the data character signals comprising a stored string. The decompressor string table is updated by storing a string having a prefix in accordance with a prior received code signal and an extension character in accordance with the first character of the currently recovered string. 6.35 A Unification 439 Here are a few other patented compression methods, some of them mentioned elsewhere in this book: 1. “Textual Substitution Data Compression with Finite Length Search Windows,” U.S. Patent 4,906,991, issued March 6, 1990 (the LZFG method). 2. “Search Tree Data Structure Encoding for Textual Substitution Data Compression Systems,” U.S. Patent 5,058,144, issued Oct. 15, 1991. The two patents above were issued to Edward Fiala and Daniel Greene. 3. “Apparatus and Method for Compressing Data Signals and Restoring the Compressed Data Signals.” This is the LZ78 patent, assigned to Sperry Corporation by the inventors Willard L. Eastman, Abraham Lempel, Jacob Ziv, and Martin Cohn. U.S. Patent 4,464,650, issued August, 1984. The following, from Ross Williams http://www.ross.net/compression/ illustrates how thorny this issue of patents is. Then, just when I thought all hope was gone, along came some software patents that drove a stake through the heart of the LZRWalgorithms by rendering them unusable. At last I was cured! I gave up compression and embarked on a new life, leaving behind the world of data compression forever. Appropriately, Dr Williams maintains a list [patents 06] of data compression-related patents. Your patent application will be denied. Your permits will be delayed. Something will force you to see reason-and to sell your drug at a lower cost. Wu had heard the argument before. And he knew Hammond was right, some new bioengineered pharmaceuticals had indeed suffered inexplicable delays and patent problems. You don’t even know exactly what you have done, but already you have reported it, patented it, and sold it. —Michael Crichton, Jurassic Park (1991) 6.35 A Unification Dictionary-based methods and methods based on prediction approach the problem of data compression from two different directions. Any method based on prediction predicts (i.e., assigns probability to) the current symbol based on its order-N context (the N symbols preceding it). Such a method normally stores many contexts of different sizes in a data structure and has to deal with frequency counts, probabilities, and probability ranges. It then uses arithmetic coding to encode the entire input stream as one large number. A dictionary-based method, on the other hand, works differently. It identifies the next phrase in the input stream, stores it in its dictionary, assigns it a code, and continues with the next phrase. Both approaches can be used to compress data because each obeys the general law of data compression, namely, to assign short codes to common events (symbols or phrases) and long codes to rare events. On the surface, the two approaches are completely different. A predictor deals with probabilities, so it can be highly efficient. At the same time, it can be expected to 440 6. Dictionary Methods be slow, since it deals with individual symbols. A dictionary-based method deals with strings of symbols (phrases), so it gobbles up the input stream faster, but it ignores correlations between phrases, typically resulting in poorer compression. The two approaches are similar because a dictionary-based method does use contexts and probabilities (although implicitly) just by storing phrases in its dictionary and searching it. The following discussion uses the LZW trie to illustrate this concept, but the argument is valid for any dictionary-based method, no matter what the details of its algorithm and its dictionary data structure. Imagine the phrase abcdef... stored in an LZW trie (Figure 6.53a). We can think of the substring abcd as the order-4 context of e. When the encoder finds another occurrence of abcde... in the input stream, it will locate our phrase in the dictionary, parse it symbol by symbol starting at the root, get to node e, and continue from there, trying to match more symbols. Eventually, the encoder will get to a leaf, where it will add another symbol and allocate another code. We can think of this process as adding a new leaf to the subtree whose root is the e of abcde.... Every time the string abcde becomes the prefix of a parse, both its subtree and its code space (the number of codes associated with it) get bigger by 1. It therefore makes sense to assign node e a probability depending on the size of its code space, and the above discussion shows that the size of the code space of node e (or, equivalently, string abcde) can be measured by counting the number of nodes of the subtree whose root is e. This is how probabilities can be assigned to nodes in any dictionary tree. The ideas of Glen Langdon in the early 1980s (see [Langdon 83] but notice that his equation (8) is wrong; it should read P(y|s) = c(s)/c(s · y); [Langdon 84] is perhaps more useful) led to a simple way of associating probabilities not just to nodes but also to edges in a dictionary tree. Assigning probabilities to edges is more useful, since the edge from node e to node f, for example, signifies an f whose context is abcde. The probability of this edge is thus the probability that an f will follow abcde in the input stream. The fact that these probabilities can be calculated in a dictionary tree shows that every dictionary-based data compression algorithm can be “simulated” by a prediction algorithm (but notice that the converse is not true). Algorithms based on prediction are, in this sense, more general, but the important fact is that these two seemingly different classes of compression methods can be unified by the observations listed here. The process whereby a dictionary encoder slides down from the root of its dictionary tree, parsing a string of symbols, can now be given a different interpretation. We can visualize it as a sequence of making predictions for individual symbols, computing codes for them, and combining the codes into one longer code, which is eventually written on the compressed stream. It is as if the code generated by a dictionary encoder for a phrase is actually made up of small chunks, each a code for one symbol. The rule for calculating the probability of the edge e › f is to count the number of nodes in the subtree whose root is f (including node f itself) and divide by the number of nodes in the subtree of e. Figure 6.53b shows a typical dictionary tree with the strings aab, baba, babc, and cac. The probabilities associated with every edge are also shown and should be easy for the reader to verify. Note that the probabilities of sibling subtrees don’t add up to 1. The probabilities of the three subtrees of the root, for example, add up to 11/12. The remaining 1/12 is assigned to the root itself and 6.35 A Unification 441 a b c d e f g Root leaf leaf (a) 3/12 5/12 3/12 2/3 2/3 1/2 1/2 1/2 3/4 4/5 1/2 a a a a a b b b c c c (b) Root 1/12 1/3 1/5 1/3 1/2 1/4 1/2 (c) Figure 6.53: Defining Probabilities in a Dictionary Tree. represents the probability that a fourth string will eventually start at the root. These “missing probabilities” are shown as horizontal lines in Figure 6.53c. The two approaches, dictionary and prediction, can be combined in a single compression method. The LZP method of Section 6.19 is one example; the LZRW4 method (Section 6.12) is another. These methods work by considering the context of a symbol before searching the dictionary. The only place where housework comes before needlework is in the dictionary. —Mary Kurtz 7 Image Compression The first part of this chapter discusses the basic features and types of digital images and the main approaches to image compression. This is followed by a description of about 30 different compression methods. We would like to start with the following observations: 1. Why were these particular methods included in the book, while others were left out? The simple answer is: Because of the documentation available to the authors. Image compression methods that are well documented were included. Methods that are proprietary, or whose documentation was not clear to the authors, were left out. 2. The treatment of the various methods is uneven. This, again, reflects the documentation available to the authors. Some methods have been documented by their developers in great detail, and this is reflected in this chapter. Where no detailed documentation was available for a compression algorithm, only its basic principles are outlined here. 3. There is no attempt to compare the various methods described here. This is because most image compression methods have been designed for a specific type of image, and also because of the practical difficulties of obtaining all the software and adapting it to run on the same platform. 4. The compression methods described in this chapter are not arranged in any particular order. After much thought and many trials, the authors gave up any hope of sorting the compression methods in any reasonable way. Readers looking for any particular method may consult the table of contents and the detailed index to easily locate it. A digital image is a rectangular array of dots, or picture elements, arranged in m rows and n columns. The expression mΧn is called the resolution of the image, and the dots are called pixels (except in the cases of fax images and video compression, where they are referred to as pels). The term “resolution” is sometimes also used to indicate the number of pixels per unit length of the image. Thus, dpi stands for dots per inch. DOI 10.1007/978-1-84882-903-9_7, © Springer-Verlag London Limited 2010 D. Salomon, G. Motta, Handbook of Data Compression, 5th ed. 444 7. Image Compression 7.1 Pixels Images are all around us. We see them in color and in high resolution. Many objects (especially artificial objects) seem perfectly smooth, with no jagged edges and no graininess. Computer graphics, on the other hand, deals with images that consist of small dots, pixels. The term pixel stands for “picture element” (see [Lyon 09] for a lively survey of the history of this important term). When we first hear of this feature of computer graphics, we tend to dismiss the entire field as trivial. It seems intuitively obvious that an image that consists of dots would always look pixelated, grainy, rough, and inferior to what we see with our eyes. Yet state-of-the-art computer-generated images are often difficult or impossible to distinguish from their real counterparts, even though they are discrete, made of pixels, and not continuous. (See also Page 526 for a discussion of human vision and the resolution of the eye.) A similar dichotomy between discrete and continuous exists in art. Many painters try to mimic nature and employ wide, continuous brush strokes to paint smooth and continuous pictures, while others choose to be pointillists. They create a painting by placing many small dots on their canvas. The most important pointillist was the 19th century French impressionist Georges Seurat. Seurat was a leader in the late 19th century neo-impressionism movement, a school of painting that uses tiny brushstrokes of contrasting colors to achieve a delicate play of light and create subtle changes in form. Seurat used this technique, which became known as pointillism or divisionism, to create large paintings made entirely of small dots of pure color. The dots are too small to be distinguished when looking at the work in its entirety, but they make his paintings shimmer with brilliance. His most well-known works are Une Baignade (1883–84) and Un dimanche apr`es-midi `a l’Ile de la Grande Jatte (1884–86). The art critic Ars`ene Alexandre had this to say about the latter painting: “Everything was so new in this immense painting—the conception was bold and the technique one that nobody had ever seen or heard before. This was the famous pointillism.” Most engineers, programmers, and users think of pixels as small squares, and this is generally true for pixels on computer monitors. Pixels in other digital output devices (displays or printers) may be rectangular or circular. However, in principle, a pixel should be considered a mathematical, dimensionless, point (see [Smith 09]). It seems impossible to reconstruct a continuous image from an array of discrete pixels, but this is precisely what the surprising Nyquist-Shannon sampling theorem [Wikipedia 03] tells us (in fact, what it guarantees). Section 10.2 (and also page 741) discuss the application of this theorem to digitized audio (a one-dimensional signal), while here we apply it to two-dimensional images. Audio is a good starting point to understand the sampling theorem. Sound fed into a microphone is converted to an electrical voltage that varies with time; it becomes a 7.1 Pixels 445 wave. A wave has a frequency, and a wave that varies all the time consists of many frequencies. We denote the maximum frequency contained in a wave by B (cycles per second, or Hertz). The samping theorem says that it is possible to reconstruct the original wave if it is sampled at a rate greater than 2B samples per second. An image is a rectilinear array of point samples (pixels). The sampling theorem guarantees that we’ll be able to reconstruct the image (i.e., to compute the color of every mathematical point in the image) if we sample the image at a rate greater than 2B pixels per unit length, where B is the maximum pixel frequency in the image. Section 7.7 explains the meaning of the term “pixel frequencies in an image,” but in practice, pixels, their values, and their frequencies depend on the accuracy of the capturing device. An ideal device should measure the color of an image at certain points, but image sensors (CCDs and CMOS) used in real devices (cameras and scanners) are often far from ideal. Because of physical limitations, manufacturing defects, and the need to capture enough light, an image sensor often measures the average color (or intensity) of a small area of the image, instead of the color at a point. Assuming that we have enough pixels for a given digital image, we compute the color of a given point in the image by interpolation. Section 9.11.7 discusses bilinear interpolation and reference [Salomon 06] discusses this and other interpolation methods and illustrates them with examples. Here, we only touch on the principles of bilinear and bicubic interpolations. The discussion assumes a grayscale image, where each pixel is a number indicating a shade of gray (an intensity), but interpolation can easily be extended to color images, where a pixel is a triplet of primary colors. Bilinear interpolation. Given enough pixels of an image and given a point Q in the image, we use bilinear interpolation to compute the intensity (grayscale) of the image at Q in the following steps: 1. We select the four pixels surrounding Q, denote their values by a, b, c, and d (Figure 7.1), and convert them to three-dimensional points by considering the value of a pixel the z coordinate of the point and adding appropriate x and y coordinates. We end up with the points P00 = (0, 0, a), P01 = (0, 1, b), P11 = (1, 1, c), and P10 = (1, 0, d). 2. We compute the parametric equations of the straight segment L1(u) running from P00 to P10 and the segment L2(u) running from P01 to P11. These equations are the simple linear interpolations L1(u) = P00(1-u)+P10u and L2(u) = P01(1-u)+P11u. 3. We compute the parametric equation Pu,w of the bilinear surface generated by interpolating segments L1(u) and L2(u). The equation is P(u,w) = L1(u)(1-w)+L2(u)w = P00(1-u)(1-w)+P10u(1-w)+P01(1-u)w+P11uw. Figure 7.1 (see also Figure 9.63) is an example of such a surface. The entire surface is spanned when parameters u and w are varied independently in the interval [0, 1]. 4. We determine the values of u and w for the given point Q and compute its intensity as the z coordinate of surface P(u,w) at these values. Bicubic interpolation. The principle of bicubic interpolation is similar to that of bilinear interpolation and is discussed in Section 7.25.7. Given a point Q on the image, we select a group of 4Χ4 pixels, convert them to three-dimensional points, and compute the bicuic surface P(u,w) that passes through these points. Once this surface is known, 446 7. Image Compression b• •c Q• a• •d            Figure 7.1: A Bilinear Surface. the values of parameters u and w for Q are determined and the intensity of image point Q becomes the height of the surface at the point determined by these values. 7.2 Image Types For the purpose of image compression it is useful to distinguish the following types of images: (For information on pixels and their history, see the lively references [Lyon 09] and [Smith 09].) 1. A bi-level (or monochromatic) image. This is an image where the pixels can have one of two values, normally referred to as black and white. Each pixel in such an image is represented by one bit, making this the simplest type of image. 2. A grayscale image. A pixel in such an image can have one of the n values 0 through n - 1, indicating one of 2n shades of gray (or shades of some other color). The value of n is normally compatible with a byte size; i.e., it is 4, 8, 12, 16, 24, or some other convenient multiple of 4 or of 8. The set of the most-significant bits of all the pixels is the most-significant bitplane. Thus, a grayscale image has n bitplanes. 3. A continuous-tone image. This type of image can have many similar colors (or grayscales). When adjacent pixels differ by just one unit, it is hard or even impossible for the eye to distinguish their colors. As a result, such an image may contain areas with colors that seem to vary continuously as the eye moves along the area. A pixel in such an image is represented by either a single large number (in the case of many grayscales) or three components (in the case of a color image). A continuous-tone image is normally a natural image (natural as opposed to artificial) and is obtained by taking a photograph with a digital camera, or by scanning a photograph or a painting. Figures 7.55 through 7.58 are typical examples of continuous-tone images. A general survey of lossless compression of this type of images is [Carpentieri et al. 00]. 7.3 Introduction 447 4. A discrete-tone image (also called a graphical image or a synthetic image). This is normally an artificial image. It may have a few colors or many colors, but it does not have the noise and blurring of a natural image. Examples are an artificial object or machine, a page of text, a chart, a cartoon, or the contents of a computer screen. (Not every artificial image is discrete-tone. A computer-generated image that’s meant to look natural is a continuous-tone image in spite of its being artificially generated.) Artificial objects, text, and line drawings have sharp, well-defined edges, and are therefore highly contrasted from the rest of the image (the background). Adjacent pixels in a discretetone image often are either identical or vary significantly in value. Such an image does not compress well with lossy methods, because the loss of just a few pixels may render a letter illegible, or change a familiar pattern to an unrecognizable one. Compression methods for continuous-tone images often do not handle sharp edges very well, so special methods are needed for efficient compression of these images. Notice that a discrete-tone image may be highly redundant, since the same character or pattern may appear many times in the image. Figure 7.59 is a typical example of a discrete-tone image. 5. A cartoon-like image. This is a color image that consists of uniform areas. Each area has a uniform color but adjacent areas may have very different colors. This feature may be exploited to obtain excellent compression. Whether an image is treated as discrete or continuous is usually dictated by the depth of the data. However, it is possible to force an image to be continuous even if it would fit in the discrete category. (From www.genaware.com) It is intuitively clear that each type of image may feature redundancy, but they are redundant in different ways. This is why any given compression method may not perform well for all images, and why different methods are needed to compress the different image types. There are compression methods for bi-level images, for continuous-tone images, and for discrete-tone images. There are also methods that try to break an image up into continuous-tone and discrete-tone parts, and compress each separately. 7.3 Introduction Modern computers employ graphics extensively. Window-based operating systems display the disk’s file directory graphically. The progress of many system operations, such as downloading a file, may also be displayed graphically. Many applications provide a graphical user interface (GUI), which makes it easier to use the program and to interpret displayed results. Computer graphics is used in many areas in everyday life to convert many types of complex information to images. Thus, images are important, but they tend to be big! Modern hardware can display many colors, which is why it is common to have a pixel represented internally as a 24-bit number, where the percentages of red, green, and blue occupy 8 bits each. Such a 24-bit pixel can specify one of 224 . 16.78 million colors. As a result, an image at a resolution of 512Χ512 that consists of such pixels occupies 786,432 bytes. At a resolution of 1024Χ1024 it becomes four times as big, requiring 3,145,728 bytes. Videos are also commonly used in computers, making for even bigger images. This is why image compression is so important. An important feature of image compression is that it can be lossy. An image, after all, exists for people 448 7. Image Compression to look at, so, when it is compressed, it is acceptable to lose image features to which the eye is not sensitive. This is one of the main ideas behind the many lossy image compression methods described in this chapter. In general, information can be compressed if it is redundant. It has been mentioned several times that data compression amounts to reducing or removing redundancy in the data. With lossy compression, however, we have a new concept, namely compressing by removing irrelevancy. An image can be lossy-compressed by removing irrelevant information even if the original image does not have any redundancy.  Exercise 7.1: It would seem that an image with no redundancy is always random (and therefore uninteresting). It that so? The idea of losing image information becomes more palatable when we consider how digital images are created. Here are three examples: (1) A real-life image may be scanned from a photograph or a painting and digitized (converted to pixels). (2) An image may be recorded by a digital camera that creates pixels and stores them directly in memory. (3) An image may be painted on the screen by means of a paint program. In all these cases, some information is lost when the image is digitized. The fact that the viewer is willing to accept this loss suggests that further loss of information might be tolerable if done properly. (Digitizing an image involves two steps: sampling and quantization. Sampling an image is the process of dividing the two-dimensional original image into small regions: pixels. Quantization is the process of assigning an integer value to each pixel. Notice that digitizing sound involves the same two steps, with the difference that sound is one-dimensional.) We present a simple process that can be employed to determine qualitatively the amount of data loss in a compressed image. Given an image A, (1) compress it to B, (2) decompress B to C, and (3) subtract D = C -A. If A was compressed without any loss and decompressed properly, then C should be identical to A and image D should be uniformly white. The more data was lost in the compression, the farther will D be from uniformly white. How should an image be compressed? The compression techniques discussed in previous chapters are RLE, scalar quantization, statistical methods, and dictionarybased methods. By itself, none is very satisfactory for color or grayscale images (although they may be used in combination with other methods). Here is why: Section 1.4.1 shows how RLE can be used for (lossless or lossy) compression of an image. This is simple, and it is used by certain parts of JPEG, especially by its lossless mode. In general, however, the other principles used by JPEG produce much better compression than does RLE alone. Facsimile compression (Section 5.7) uses RLE combined with Huffman coding and obtains good results, but only for bi-level images. Scalar quantization has been mentioned in Section 1.6. It can be used to compress images, but its performance is mediocre. Imagine an image with 8-bit pixels. It can be compressed with scalar quantization by cutting off the four least-significant bits of each pixel. This yields a compression ratio of 0.5, not very impressive, and at the same time reduces the number of colors (or grayscales) from 256 to just 16. Such a reduction not only degrades the overall quality of the reconstructed image, but may also create bands of different colors, a noticeable and annoying effect that’s illustrated here. 7.3 Introduction 449 Imagine a row of 12 pixels with similar colors, ranging from 202 to 215. In binary notation these values are 11010111 11010110 11010101 11010011 11010010 11010001 11001111 11001110 11001101 11001100 11001011 11001010. Quantization will result in the 12 4-bit values 1101 1101 1101 1101 1101 1101 1100 1100 1100 1100 1100 1100, which will reconstruct the 12 pixels 11010000 11010000 11010000 11010000 11010000 11010000 11000000 11000000 11000000 11000000 11000000 11000000. The first six pixels of the row now have the value 110100002 = 208, while the next six pixels are 110000002 = 192. If neighboring rows have similar pixels, the first six columns will form a band, distinctly different from the band formed by the next six columns. This banding, or contouring, effect is very noticeable to the eye, since our eyes are sensitive to edges and breaks in an image. One way to eliminate this effect is called improved grayscale (IGS) quantization. It works by adding to each pixel a random number generated from the four rightmost bits of previous pixels. Section 7.4.1 shows that the least-significant bits of a pixel are fairly random, so IGS works by adding to each pixel randomness that depends on the neighborhood of the pixel. The method maintains an 8-bit variable, denoted by rsm, that’s initially set to zero. For each 8-bit pixel P to be quantized (except the first one), the IGS method does the following: 1. Set rsm to the sum of the eight bits of P and the four rightmost bits of rsm. However, if P has the form 1111xxxx, set rsm to P. 2. Write the four leftmost bits of rsm on the compressed stream. This is the compressed value of P. IGS is thus not exactly a quantization method, but a variation of scalar quantization. The first pixel is quantized in the usual way, by dropping its four rightmost bits. Table 7.2 illustrates the operation of IGS. Compressed Pixel Value rsm value 1 1010 0110 0000 0000 1010 2 1101 0010 1101 0010 1101 3 1011 0101 1011 0111 1011 4 1001 1100 1010 0011 1010 5 1111 0100 1111 0100 1111 6 1011 0011 1011 0111 1011 Table 7.2: Illustrating the IGS Method. Vector quantization can be used more successfully to compress images. It is discussed in Section 7.19. Statistical methods work best when the symbols being compressed have different probabilities. An input stream where all symbols have the same probability will not compress, even though it may not be random. It turns out that in a continuous-tone color or grayscale image, the different colors or shades of gray may often have roughly the same probabilities. This is why statistical methods are not a good choice for compressing such 450 7. Image Compression images, and why new approaches are needed. Images with color discontinuities, where adjacent pixels have widely different colors, compress better with statistical methods, but it is not easy to predict, just by looking at an image, whether it has enough color discontinuities. (a) (b) (c) An ideal vertical rule is shown in (a). In (b), the rule is assumed to be perfectly digitized into ten pixels, stacked vertically. However, if the image is placed in the scanner slightly crooked, the scanning may be imperfect, and the resulting pixels might look as in (c). Figure 7.3: Perfect and Imperfect Digitizing. Dictionary-based compression methods also tend to be unsuccessful in dealing with continuous-tone images. Such an image typically contains adjacent pixels with similar colors, but does not contain repeating patterns. Even an image that contains repeated patterns such as vertical lines may lose them when digitized. A vertical line in the original image may become slightly crooked when the image is digitized (Figure 7.3), so the pixels in a scan row may end up having slightly different colors from those in neighboring rows, resulting in a dictionary with short strings. (This problem may also affect curved edges.) Another problem with dictionary compression of images is that such methods scan the image row by row, and therefore may miss vertical correlations between pixels. An example is the two simple images of Figure 7.4a,b. Saving both in GIF89, a dictionarybased graphics file format (Section 6.21), has resulted in file sizes of 1053 and 1527 bytes, respectively, on the author’s computer. (a) (b) Figure 7.4: Dictionary Compression of Parallel Lines. 7.3 Introduction 451 Traditional methods are therefore unsatisfactory for image compression, so this chapter discusses novel approaches. They are all different, but they remove redundancy from an image by using the following principle (see also Section 1.4): The Principle of Image Compression. If we select a pixel in an image at random, there is a good chance that its neighbors will have the same color or very similar colors. Image compression is therefore based on the fact that neighboring pixels are highly correlated. This correlation is also called spatial redundancy. Here is a simple example that illustrates what can be done with correlated pixels. The following sequence of values gives the intensities of 24 adjacent pixels in a row of a continuous-tone image: 12, 17, 14, 19, 21, 26, 23, 29, 41, 38, 31, 44, 46, 57, 53, 50, 60, 58, 55, 54, 52, 51, 56, 60. Only two of the 24 pixels are identical. Their average value is 40.3. Subtracting pairs of adjacent pixels results in the sequence 12, 5, -3, 5, 2, 4, -3, 6, 11, -3, -7, 13, 4, 11, -4, -3, 10, -2, -3, 1, -2, -1, 5, 4. The two sequences are illustrated in Figure 7.5. Figure 7.5: Values and Differences of 24 Adjacent Pixels. The sequence of difference values has three properties that illustrate its compression potential: (1) The difference values are smaller than the original pixel values. Their average is 2.58. (2) They repeat. There are just 15 distinct difference values, so in principle they can be coded by four bits each. (3) They are decorrelated: adjacent difference values tend to be different. This can be seen by subtracting them, which results in the sequence of 24 second differences 12, -7, -8, 8, -3, 2, -7, 9, 5, -14, -4, 20, -11, 7, -15, 1, 13, -12, -1, 4, -3, 1, 6, 1. They are larger than the differences themselves. Figure 7.6 provides another illustration of the meaning of the words “correlated quantities.” A 32 Χ 32 matrix A is constructed of random numbers, and its elements are displayed in part (a) as shaded squares. The random nature of the elements is clear. The matrix is then inverted and stored in B, which is shown in part (b). This time, there seems to be more structure to the 32 Χ 32 squares. A direct calculation using Equation (7.1) shows that the cross-correlation between the top two rows of A is 0.0412, 452 7. Image Compression whereas the cross-correlation between the top two rows of B is -0.9831. The elements of B are correlated since each depends on all the elements of A R = nxiyi -xiyi [nx2i - (xi)2][ny2 i - (yi)2] . (7.1) 5 10 15 20 25 30 5 10 15 20 25 30 5 10 15 20 25 30 5 10 15 20 25 30 (a) (b) n=32; a=rand(n); imagesc(a); colormap(gray) b=inv(a); imagesc(b) Figure 7.6: Maps of (a) a Random Matrix and (b) its Inverse.  Exercise 7.2: Use mathematical software to illustrate the covariance matrices of (1) a matrix with correlated values and (2) a matrix with decorrelated values. Once the concept of correlated quantities is grasped, we start looking for a correlation test. Given a matrix M, a statistical test is needed to determine whether its elements are correlated or not. The test is based on the statistical concept of covariance. If the elements of M are decorrelated (i.e., independent), then the covariance of any two different rows and any two different columns of M will be zero (the covariance of a row or of a column with itself is always 1). As a result, the covariance matrix of M (whether covariance of rows or of columns) will be diagonal. If the covariance matrix of M is not diagonal, then the elements of M are correlated. The statistical concepts of variance, covariance, and correlation are discussed in any text on statistics. The principle of image compression has another aspect. We know from experience that the brightness of neighboring pixels is also correlated. Two adjacent pixels may have different colors. One may be mostly red, and the other may be mostly green. Yet if the red component of the first is bright, the green component of its neighbor will, in most cases, also be bright. This property can be exploited by converting pixel representations from RGB to three other components, one of which is the brightness, and the other two represent color. One such format (or color space) is YCbCr, where Y (the “luminance” component) represents the brightness of a pixel, and Cb and Cr 7.4 Approaches to Image Compression 453 define its color. This format is discussed in Section 7.10.1, but its advantage is easy to understand. The eye is sensitive to small changes in brightness but not to small changes in color. Thus, losing information in the Cb and Cr components compresses the image while introducing distortions to which the eye is not sensitive. Losing information in the Y component, on the other hand, is very noticeable to the eye. 7.4 Approaches to Image Compression An image compression method is normally designed for a specific type of image, and this section lists various approaches to compressing images of different types. Only the general principles are discussed here; specific methods are described in the remainder of this chapter. Approach 1: This is appropriate for bi-level images. A pixel in such an image is represented by one bit. Applying the principle of image compression to a bi-level image therefore means that the immediate neighbors of a pixel P tend to be identical to P. Thus, it makes sense to use run-length encoding (RLE) to compress such an image. A compression method for such an image may scan it in raster order (row by row) and compute the lengths of runs of black and white pixels. The lengths are encoded by variable-length codes and are written on the compressed stream. An example of such a method is facsimile compression, Section 5.7. It should be stressed that this is just an approach to bi-level image compression. The details of specific methods vary. For instance, a method may scan the image column by column or in zigzag (Figure 1.8b), it may convert the image to a quadtree (Section 7.34), or it may scan it region by region using a space-filling curve (Section 7.36). Approach 2: Also for bi-level images. The principle of image compression tells us that the neighbors of a pixel tend to be similar to the pixel. We can extend this principle and conclude that if the current pixel has color c (where c is either black or white), then pixels of the same color seen in the past (and also those that will be found in the future) tend to have the same immediate neighbors. This approach looks at n of the near neighbors of the current pixel and considers them an n-bit number. This number is the context of the pixel. In principle there can be 2n contexts, but because of image redundancy we expect them to be distributed in a nonuniform way. Some contexts should be common while others will be rare. The encoder counts how many times each context has already been found for a pixel of color c, and assigns probabilities to the contexts accordingly. If the current pixel has color c and its context has probability p, the encoder can use adaptive arithmetic coding to encode the pixel with that probability. This approach is used by JBIG (Section 7.14). Next, we turn to grayscale images. A pixel in such an image is represented by n bits and can have one of 2n values. Applying the principle of image compression to a grayscale image implies that the immediate neighbors of a pixel P tend to be similar to P, but are not necessarily identical. Thus, RLE should not be used to compress such an image. Instead, two approaches are discussed. Approach 3: Separate the grayscale image into n bi-level images and compress each with RLE and prefix codes. The principle of image compression seems to imply intuitively that two adjacent pixels that are similar in the grayscale image will be identical 454 7. Image Compression in most of the n bi-level images. This, however, is not true, as the following example makes clear. Imagine a grayscale image with n = 4 (i.e., 4-bit pixels, or 16 shades of gray). The image can be separated into four bi-level images. If two adjacent pixels in the original grayscale image have values 0000 and 0001, then they are similar. They are also identical in three of the four bi-level images. However, two adjacent pixels with values 0111 and 1000 are also similar in the grayscale image (their values are 7 and 8, respectively) but differ in all four bi-level images. This problem occurs because the binary codes of adjacent integers may differ by several bits. The binary codes of 0 and 1 differ by one bit, those of 1 and 2 differ by two bits, and those of 7 and 8 differ by four bits. The solution is to design special binary codes such that the codes of any consecutive integers i and i + 1 will differ by one bit only. An example of such a code is the reflected Gray codes of Section 7.4.1. Approach 4: Use the context of a pixel to predict its value. The context of a pixel is the values of some of its neighbors. We can examine some neighbors of a pixel P, compute an average A of their values, and predict that P will have the value A. The principle of image compression tells us that our prediction will be correct in most cases, almost correct in many cases, and completely wrong in a few cases. We can say that the predicted value of pixel P represents the redundant information in P. We now calculate the difference . def = P - A, and assign variable-length codes to the different values of . such that small values (which we expect to be common) are assigned short codes and large values (which are expected to be rare) are assigned long codes. If P can have the values 0 through m-1, then values of . are in the range [-(m-1), +(m-1)], and the number of codes needed is 2(m- 1) + 1 or 2m- 1. Experiments with a large number of images suggest that the values of . tend to be distributed according to the Laplace distribution (Figure 7.143b). A compression method can, therefore, use this distribution to assign a probability to each value of ., and use arithmetic coding to encode the . values very efficiently. This is the principle of the MLP method (Section 7.25). The context of a pixel may consist of just one or two of its immediate neighbors. However, better results may be obtained when several neighbor pixels are included in the context. The average A in such a case should be weighted, with near neighbors assigned higher weights (see, for example, Table 7.141). Another important consideration is the decoder. In order for it to decode the image, it should be able to compute the context of every pixel. This means that the context should employ only pixels that have already been encoded. If the image is scanned in raster order, the context should include only pixels located above the current pixel or on the same row and to its left. Approach 5: Transform the values of the pixels and encode the transformed values. The concept of a transform, as well as the most important transforms used in image compression, are discussed in Section 7.6. Chapter 8 is devoted to the wavelet transform. Recall that compression is achieved by reducing or removing redundancy. The redundancy of an image is caused by the correlation between pixels, so transforming the pixels to a representation where they are decorrelated eliminates the redundancy. It is also possible to think of a transform in terms of the entropy of the image. In a highly 7.4 Approaches to Image Compression 455 correlated image, the pixels tend to have equiprobable values, which results in maximum entropy. If the transformed pixels are decorrelated, certain pixel values become common, thereby having large probabilities, while others are rare. This results in small entropy. Quantizing the transformed values can produce efficient lossy image compression. We want the transformed values to be independent because coding independent values makes it simpler to construct a statistical model. We now turn to color images. A pixel in such an image consists of three color components, such as red, green, and blue. Most color images are either continuous-tone or discrete-tone. Approach 6: The principle of this approach is to separate a continuous-tone color image into three grayscale images and compress each of the three separately, using approaches 3, 4, or 5. For a continuous-tone image, the principle of image compression implies that adjacent pixels have similar, although perhaps not identical, colors. However, similar colors do not mean similar pixel values. Consider, for example, 12-bit pixel values where each color component is expressed in four bits. Thus, the 12 bits 1000|0100|0000 represent a pixel whose color is a mixture of eight units of red (about 50%, since the maximum is 15 units), four units of green (about 25%), and no blue. Now imagine two adjacent pixels with values 0011|0101|0011 and 0010|0101|0011. They have similar colors, since only their red components differ, and only by one unit. However, when considered as 12-bit numbers, the two numbers 001101010011 and 001001010011 are very different, since they differ in one of their most significant bits. An important feature of this approach is to use a luminance chrominance color representation instead of the more common RGB. The concepts of luminance and chrominance are discussed in Section 7.10.1 and in [Salomon 99]. The advantage of the luminance chrominance color representation is that the eye is sensitive to small changes in luminance but not in chrominance. This allows the loss of considerable data in the chrominance components, while making it possible to decode the image without a significant visible loss of quality. Approach 7: A different approach is needed for discrete-tone images. Recall that such an image contains uniform regions, and a region may appear several times in the image. A good example is a screen dump. Such an image consists of text and icons. Each character of text and each icon is a region, and any region may appear several times in the image. A possible way to compress such an image is to scan it, identify regions, and find repeating regions. If a region B is identical to an already found region A, then B can be compressed by writing a pointer to A on the compressed stream. The block decomposition method (FABD, Section 7.32) is an example of how this approach can be implemented. Approach 8: Partition the image into parts (overlapping or not) and compress it by processing the parts one by one. Suppose that the next unprocessed image part is part number 15. Try to match it with parts 1–14 that have already been processed. If part 15 can be expressed, for example, as a combination of parts 5 (scaled) and 11 (rotated), then only the few numbers that specify the combination need be saved, and part 15 can be discarded. If part 15 cannot be expressed as a combination of already-processed parts, it is declared processed and is saved in raw format. 456 7. Image Compression This approach is the basis of the various fractal methods for image compression. It applies the principle of image compression to image parts instead of to individual pixels. Applied this way, the principle tells us that “interesting” images (i.e., those that are being compressed in practice) have a certain amount of self similarity. Parts of the image are identical or similar to the entire image or to other parts. Image compression methods are not limited to these basic approaches. This book discusses methods that use the concepts of context trees, Markov models (Section 11.8), and wavelets, among others. In addition, the concept of progressive image compression (Section 7.13) should be mentioned, since it adds another dimension to the field of image compression. 7.4.1 Gray Codes An image compression method that has been developed specifically for a certain type of image can sometimes be used for other types. Any method for compressing bilevel images, for example, can be used to compress grayscale images by separating the bitplanes and compressing each individually, as if it were a bi-level image. Imagine, for example, an image with 16 grayscale values. Each pixel is specified by four bits, so the image can be separated into four bi-level images. The trouble with this approach is that it violates the general principle of image compression. Imagine two adjacent 4-bit pixels with values 7 = 01112 and 8 = 10002. These pixels have close values, but when separated into four bitplanes, the resulting 1-bit pixels are different in every bitplane! This is because the binary representations of the consecutive integers 7 and 8 differ in all four bit positions. In order to apply any bi-level compression method to grayscale images, a binary representation of the integers is needed where consecutive integers have codes differing by one bit only. Such a representation exists and is called reflected Gray code (RGC). This code is easy to generate with the following recursive construction: Start with the two 1-bit codes (0, 1). Construct two sets of 2-bit codes by duplicating (0, 1) and appending, either on the left or on the right, first a zero, then a one, to the original set. The result is (00, 01) and (10, 11). We now reverse (reflect) the second set, and concatenate the two. The result is the 2-bit RGC (00, 01, 11, 10); a binary code of the integers 0 through 3 where consecutive codes differ by exactly one bit. Applying the rule again produces the two sets (000, 001, 011, 010) and (110, 111, 101, 100), which are concatenated to form the 3-bit RGC. Note that the first and last codes of any RGC also differ by one bit. Here are the first three steps for computing the 4-bit RGC: Add a zero (0000, 0001, 0011, 0010, 0110, 0111, 0101, 0100), Add a one (1000, 1001, 1011, 1010, 1110, 1111, 1101, 1100), reflect (1100, 1101, 1111, 1110, 1010, 1011, 1001, 1000). Table 7.7 shows how individual bits change when moving through the binary codes of the first 32 integers. The 5-bit binary codes of these integers are listed in the oddnumbered columns of the table, with the bits of integer i that differ from those of i - 1 shown in boldface. It is easy to see that the least-significant bit (bit b0) changes all the time, bit b1 changes for every other number, and, in general, bit bk changes every k integers. The even-numbered columns list one of the several possible reflected Gray codes for these integers. A recursive Matlab function to compute RGC is also listed. 7.4 Approaches to Image Compression 457 43210 Gray 43210 Gray 43210 Gray 43210 Gray 00000 00000 01000 10010 10000 00011 11000 10001 00001 00100 01001 10110 10001 00111 11001 10101 00010 01100 01010 11110 10010 01111 11010 11101 00011 01000 01011 11010 10011 01011 11011 11001 00100 11000 01100 01010 10100 11011 11100 01001 00101 11100 01101 01110 10101 11111 11101 01101 00110 10100 01110 00110 10110 10111 11110 00101 00111 10000 01111 00010 10111 10011 11111 00001 function b=rgc(a,i) [r,c]=size(a); b=[zeros(r,1),a; ones(r,1),flipud(a)]; if i>1, b=rgc(b,i-1); end; Table 7.7: First 32 Binary and Reflected Gray Codes.  Exercise 7.3: It is also possible to generate the reflected Gray code of an integer n with the following nonrecursive rule: Exclusive-OR n with a copy of itself that’s logically shifted one position to the right. In the C programming language this is denoted by n^(n>>1). Use this expression to construct a table similar to Table 7.7. clear; clear; filename=’parrots128’; dim=128; filename=’parrots128’; dim=128; fid=fopen(filename,’r’); fid=fopen(filename,’r’); img=fread(fid,[dim,dim])’; img=fread(fid,[dim,dim])’; mask=1; % between 1 and 8 mask=1 % between 1 and 8 a=bitshift(img,-1); b=bitxor(img,a); nimg=bitget(img,mask); nimg=bitget(b,mask); imagesc(nimg), colormap(gray) imagesc(nimg), colormap(gray) Binary code Gray code Figure 7.8: Matlab Code to Separate Image Bitplanes. The conclusion is that the most-significant bitplanes of an image obey the principle of image compression more than the least-significant ones. When adjacent pixels have values that differ by one unit (such as p and p + 1), chances are that the leastsignificant bits are different and the most-significant ones are identical. Any image compression method that compresses bitplanes individually should therefore treat the least-significant bitplanes differently from the most-significant ones, or should use RGC instead of the binary code to represent pixels. Figures 7.10, 7.11, and 7.12 (prepared by the Matlab code of Figure 7.8) show the eight bitplanes of the well-known parrots 458 7. Image Compression image in both the binary code (the left column) and in RGC (the right column). The bitplanes are numbered 8 (the leftmost or most-significant bits) through 1 (the rightmost or least-significant bits). It is obvious that the least-significant bitplane doesn’t show any correlations between the pixels; it is random or very close to random in both binary and RGC. Bitplanes 2 through 5, however, exhibit better pixel correlation in the Gray code. Bitplanes 6 through 8 look different in Gray code and binary, but seem to be highly correlated in either representation. 43210 Gray 43210 Gray 43210 Gray 43210 Gray 00000 00000 01000 01100 10000 11000 11000 10100 00001 00001 01001 01101 10001 11001 11001 10101 00010 00011 01010 01111 10010 11011 11010 10111 00011 00010 01011 01110 10011 11010 11011 10110 00100 00110 01100 01010 10100 11110 11100 10010 00101 00111 01101 01011 10101 11111 11101 10011 00110 00101 01110 01001 10110 11101 11110 10001 00111 00100 01111 01000 10111 11100 11111 10000 a=linspace(0,31,32); b=bitshift(a,-1); b=bitxor(a,b); dec2bin(b) Table 7.9: First 32 Binary and Gray Codes. Figure 7.13 is a graphic representation of two versions of the first 32 reflected Gray codes. Part (b) shows the codes of Table 7.7, and part (c) shows the codes of Table 7.9. Even though both are Gray codes, they differ in the way the bits in each bitplane alternate between 0 and 1. In part (b), the bits of the most-significant bitplane alternate four times between 0 and 1. Those of the second most-significant bitplane alternate eight times between 0 and 1, and the bits of the remaining three bitplanes alternate 16, two, and one times between 0 and 1. When the bitplanes are separated, the middle bitplane features the smallest correlation between the pixels, since the Gray codes of adjacent integers tend to have different bits in this bitplane. The Gray codes shown in Figure 7.13c, on the other hand, alternate more and more between 0 and 1 as we move from the most significant bitplanes to the least-significant ones. The least significant bitplanes of this version feature less and less correlation between the pixels and therefore tend to be random. For comparison, Figure 7.13a shows the binary code. It is obvious that bits in this code alternate more often between 0 and 1.  Exercise 7.4: Even a cursory look at the Gray codes of Figure 7.13c shows that they exhibit some regularity. Examine these codes carefully and identify two features that may be used to compute the codes.  Exercise 7.5: Figure 7.13 is a graphic representation of the binary codes and reflected Gray codes. Find a similar graphic representation of the same codes that illustrates the fact that the first and last codes also differ by one bit. 7.4 Approaches to Image Compression 459 (1) (2) Binary code Gray code Figure 7.10: Bitplanes 1,2 of the Parrots Image. 460 7. Image Compression (3) (4) (5) Binary code Gray code Figure 7.11: Bitplanes 3, 4, and 5 of the Parrots Image. 7.4 Approaches to Image Compression 461 (6) (7) (8) Binary code Gray code Figure 7.12: Bitplanes 6, 7, and 8 of the Parrots Image. 462 7. Image Compression b4 b3 b2 b1 b0 0123456789 10 11 12 13 14 15 16 17 18 19 21 21 22 23 24 25 26 27 28 29 30 31 1 3 7 15 31 b4 b3 b2 b1 b0 0123456789 10 11 12 13 14 15 16 17 18 19 21 21 22 23 24 25 26 27 28 29 30 31 4 8 16 2 1 b4 b3 b2 b1 b0 0123456789 10 11 12 13 14 15 16 17 18 19 21 21 22 23 24 25 26 27 28 29 30 31 1 2 4 8 16 (a) (b) (c) Table 7.13: First 32 Binary and Reflected Gray Codes. The binary Gray code is fun, For in it strange things can be done. Fifteen, as you know, Is one, oh, oh, oh, And ten is one, one, one, one. —Anonymous 7.4 Approaches to Image Compression 463 Color images provide another example of using the same compression method across image types. Any compression method for grayscale images can be used to compress color images. In a color image, each pixel is represented by three color components (such as RGB). Imagine a color image where each color component is represented by one byte. A pixel is represented by three bytes, or 24 bits, but these bits should not be considered a single number. The two pixels 118|206|12 and 117|206|12 differ by just one unit in the first component, so they have very similar colors. Considered as 24-bit numbers, however, these pixels are very different, since they differ in one of their mostsignificant bits. Any compression method that treats these pixels as 24-bit numbers would consider these pixels very different, and its performance would suffer as a result. A compression method for grayscale images can be applied to compressing color images, but the color image should first be separated into three color components, and each component compressed individually as a grayscale image. For an example of the use of RGC for image compression see Section 7.31. History of Gray Codes Gray codes are named after Frank Gray, who patented their use for shaft encoders in 1953 [Gray 53]. However, the work was performed much earlier, the patent being applied for in 1947. Gray was a researcher at Bell Telephone Laboratories. During the 1930s and 1940s he was awarded numerous patents for work related to television. According to [Heath 72] the code was first, in fact, used by J. M. E. Baudot for telegraphy in the 1870s (Section 1.1.4), though it is only since the advent of computers that the code has become widely known. The Baudot code uses five bits per symbol. It can represent 32 Χ 2 - 2 = 62 characters (each code can have two meanings, the meaning being indicated by the LS and FS codes). It became popular and, by 1950, was designated the International Telegraph Code No. 1. It was used by many first- and second-generation computers. The August 1972 issue of Scientific American contains two articles of interest, one on the origin of binary codes [Heath 72], and another [Gardner 72] on some entertaining aspects of the Gray codes. 7.4.2 Error Metrics Developers and implementers of lossy image compression methods need a standard metric to measure the quality of reconstructed images compared with the original ones. The better a reconstructed image resembles the original one, the bigger should be the value produced by this metric. Such a metric should also produce a dimensionless number, and that number should not be very sensitive to small variations in the reconstructed image. A common measure used for this purpose is the peak signal to noise ratio (PSNR). It is familiar to workers in the field, it is also simple to calculate, but it has only a limited, approximate relationship with the perceived errors noticed by the human visual system. This is why higher PSNR values imply closer resemblance between the reconstructed and the original images, but they do not provide a guarantee that viewers will like the reconstructed image. Denoting the pixels of the original image by Pi and the pixels of the reconstructed image by Qi (where 1 . i . n), we first define the mean square error (MSE) between 464 7. Image Compression the two images as MSE = 1 n n i=1 (Pi - Qi)2. (7.2) It is the average of the square of the errors (pixel differences) of the two images. The root mean square error (RMSE) is defined as the square root of the MSE, and the PSNR is defined as PSNR = 20 log10 maxi |Pi| RMSE , (7.3) (but see Equation (9.4) for a different version). The absolute value is normally not needed, since pixel values are rarely negative. For a bi-level image, the numerator is 1. For a grayscale image with eight bits per pixel, the numerator is 255. For color images, only the luminance component is used. Greater resemblance between the images implies smaller RMSE and, as a result, larger PSNR. The PSNR is dimensionless, since the units of both numerator and denominator are pixel values. However, because of the use of the logarithm, we say that the PSNR is expressed in decibels (dB, Section 10.1). The use of the logarithm also implies less sensitivity to changes in the RMSE. For example, dividing the RMSE by 10 multiplies the PSNR by 2. Notice that the PSNR has no absolute meaning. It is meaningless to say that a PSNR of, say, 25 is good. PSNR values are used only to compare the performance of different lossy compression methods or the effects of different parametric values on the performance of an algorithm. The MPEG committee, for example, uses an informal threshold of PSNR = 0.5 dB to decide whether to incorporate a coding optimization, since they believe that an improvement of that magnitude would be visible to the eye. Typical PSNR values range between 20 and 40. Assuming pixel values in the range [0, 255], an RMSE of 25.5 results in a PSNR of 20, and an RMSE of 2.55 results in a PSNR of 40. An RMSE of zero (i.e., identical images) results in an infinite (or, more precisely, undefined) PSNR. An RMSE of 255 results in a PSNR of zero, and RMSE values greater than 255 yield negative PSNRs.  Exercise 7.6: If the maximum pixel value is 255, can the RMSE values be greater than 255? Some authors define the PSNR as PSNR = 10 log10 maxi |Pi|2 MSE . In order for the two formulations to produce the same result, the logarithm is multiplied in this case by 10 instead of by 20, since log10 A2 = 2 log10 A. Either definition is useful, because only relative PSNR values are used in practice. However, the use of two different factors is confusing. A related measure is signal to noise ratio (SNR). This is defined as SNR = 20 log10 )1 n n i=1 P2 i RMSE . 7.4 Approaches to Image Compression 465 The numerator is the root mean square of the original image. Figure 7.14 is a Matlab function to compute the PSNR of two images. A typical call is PSNR(A,B), where A and B are image files. They must have the same resolution and have pixel values in the range [0, 1]. function PSNR(A,B) if A==B error(’Images are identical; PSNR is undefined’) end max2_A=max(max(A)); max2_B=max(max(B)); min2_A=min(min(A)); min2_B=min(min(B)); if max2_A>1 | max2_B>1 | min2_A<0 | min2_B<0 error(’pixels must be in [0,1]’) end differ=A-B; decib=20*log10(1/(sqrt(mean(mean(differ.^2))))); disp(sprintf(’PSNR = +%5.2f dB’,decib)) Figure 7.14: A Matlab Function to Compute PSNR. Another relative of the PSNR is the signal to quantization noise ratio (SQNR) This is a measure of the effect of quantization on signal quality. It is defined as SQNR = 10 log10 signal power quantization error , where the quantization error is the difference between the quantized signal and the original signal. Another approach to the comparison of an original and a reconstructed image is to generate the difference image and judge it visually. Intuitively, the difference image is Di = Pi-Qi, but such an image is hard to judge visually because its pixel values Di tend to be small numbers. If a pixel value of zero represents white, such a difference image would be almost invisible. In the opposite case, where pixel values of zero represent black, such a difference would be too dark to judge. Better results are obtained by calculating Di = a(Pi - Qi) + b, where a is a magnification parameter (typically a small number such as 2) and b is half the maximum value of a pixel (typically 128). Parameter a serves to magnify small differences, while b shifts the difference image from extreme white (or extreme black) to a more comfortable gray. 466 7. Image Compression 7.5 Intuitive Methods It is easy to come up with simple, intuitive methods for compressing images. They are inefficient and are described here for the sake of completeness. 7.5.1 Subsampling Subsampling is perhaps the simplest way to compress an image. One approach to subsampling is simply to delete some of the pixels. The encoder may, for example, ignore every other row and every other column of the image, and write the remaining pixels (which constitute 25% of the image) on the compressed stream. The decoder inputs the compressed data and uses each pixel to generate four identical pixels of the reconstructed image. This, of course, involves the loss of much image detail and is rarely acceptable. Notice that the compression ratio is known in advance. A slight improvement is obtained when the encoder calculates the average of each block of four pixels and writes this average on the compressed stream. No pixel is totally deleted, but the method is still primitive, because a good lossy image compression method should lose only data to which the eye is not sensitive. Better results (but worse compression) are obtained when the color representation of the image is changed from the original (normally RGB) to luminance and chrominance. The encoder subsamples the two chrominance components of a pixel but not its luminance component. Assuming that each component uses the same number of bits, the two chrominance components use 2/3 of the image size. Subsampling them reduces this to 25% of 2/3, or 1/6. The size of the compressed image is therefore 1/3 (for the uncompressed luminance component), plus 1/6 (for the two chrominance components) or 1/2 of the original size. 7.5.2 Quantization Scalar quantization has been mentioned in Section 7.3. This is an intuitive, lossy method where the information lost is not necessarily the least important. Vector quantization can obtain better results, and an intuitive version of it is described here. The image is partitioned into equal-size blocks (called vectors) of pixels, and the encoder has a list (called a codebook) of blocks of the same size. Each image block B is compared to all the blocks of the codebook and is matched with the “closest” one. If B is matched with codebook block C, then the encoder writes a pointer to C on the compressed stream. If the pointer is smaller than the block size, compression is achieved. Figure 7.15 shows an example. The details of selecting and maintaining the codebook and of matching blocks are discussed in Section 7.19. Notice that vector quantization is a method where the compression ratio is known in advance. 7.6 Image Transforms 467 Codebook Original image Compressed image Reconstructed image 0 0 0 1 2 3 3 2 4 4 5 6 6 7 Figure 7.15: Intuitive Vector Quantization. 7.6 Image Transforms The mathematical concept of a transform is a powerful tool that is employed in many areas and can also serve as an approach to image compression. Section 8.1 discusses this concept in general, as well as the Fourier transform. An image can be compressed by transforming its pixels (which are correlated) to a representation where they are decorrelated. Compression is achieved if the new values are smaller, on average, than the original ones. Lossy compression can be achieved by quantizing the transformed values. The decoder inputs the transformed values from the compressed stream and reconstructs the (precise or approximate) original data by applying the inverse transform. The transforms discussed in this section are orthogonal. Section 8.6.1 discusses subband transforms. One of several excellent references on transforms and their applications to data compression is [Rao and Yip 00]. The term decorrelated means that the transformed values are independent of one another. As a result, they can be encoded independently, which makes it simpler to construct a statistical model. An image can be compressed if its representation has redundancy. The redundancy in images stems from pixel correlation. If we transform the image to a representation where the pixels are decorrelated, we have eliminated the redundancy and the image has been fully compressed. We start with a simple example, where we scan an image in raster order and group pairs of adjacent pixels. Because the pixels are correlated, the two pixels (x, y) of a pair normally have similar values. We now consider each pair of pixels a point in twodimensional space, and we plot the points. We know that all the points of the form (x, x) are located on the 45. line y = x, so we expect our points to be concentrated around this line. Figure 7.17a shows the results of plotting the pixels of a typical image—where a pixel has values in the interval [0, 255]—in such a way. Most points form a cloud around this line, and only a few points are located away from it. We now transform the image by rotating all the points 45. clockwise about the origin, such that the 45. line now coincides with the x-axis (Figure 7.17b). This is done by the simple transformation [see Equation (7.63)] (x*, y*) = (x, y)cos 45. -sin 45. sin 45. cos 45. = (x, y) 1 .2 1 -1 1 1 = (x, y)R, (7.4) 468 7. Image Compression where the rotation matrix R is orthonormal (i.e., the dot product of a row with itself is 1, the dot product of different rows is 0, and the same is true for columns). The inverse transformation is (x, y) = (x*, y*)R-1 = (x*, y*)RT = (x*, y*) 1 .2  1 1 -1 1 . (7.5) (The inverse of an orthonormal matrix is its transpose.) It is obvious that most points end up with y coordinates that are zero or close to zero, while the x coordinates don’t change much. Figure 7.18a,b shows the distributions of the x and y coordinates (i.e., the odd-numbered and even-numbered pixels) of the 128Χ 128 Χ 8 grayscale Lena image before the rotation. It is clear that the two distributions don’t differ by much. Figure 7.18c,d shows that the distribution of the x coordinates stays about the same (with greater variance) but the y coordinates are concentrated around zero. The Matlab code that generated these results is also shown. (Figure 7.18d shows that the y coordinates are concentrated around 100, but this is because a few were as small as -101, so they had to be scaled by 101 to fit in a Matlab array, which always starts at index 1.) p={{5,5},{6, 7},{12.1,13.2},{23,25},{32,29}}; rot={{0.7071,-0.7071},{0.7071,0.7071}}; Sum[p[[i,1]]p[[i,2]], {i,5}] q=p.rot Sum[q[[i,1]]q[[i,2]], {i,5}] Figure 7.16: Code for Rotating Five Points. Once the coordinates of points are known before and after the rotation, it is easy to measure the reduction in correlation. A simple measure is the sum i xiyi, also called the cross-correlation of points (xi, yi).  Exercise 7.7: Given the five points (5, 5), (6, 7), (12.1, 13.2), (23, 25), and (32, 29), rotate them 45. clockwise and calculate their cross-correlations before and after the rotation. We can now compress the image by simply writing the transformed pixels on the compressed stream. If lossy compression is acceptable, then all the pixels can be quantized (Sections 1.6 and 7.19), resulting in even smaller numbers. We can also write all the odd-numbered pixels (those that make up the x coordinates of the pairs) on the compressed stream, followed by all the even-numbered pixels. These two sequences are called the coefficient vectors of the transform. The latter sequence consists of small numbers and may, after quantization, have runs of zeros, resulting in even better compression. It can be shown that the total variance of the pixels does not change by the rotation, because a rotation matrix is orthonormal. However, since the variance of the new y coordinates is small, most of the variance is now concentrated in the x coordinates. The variance is sometimes called the energy of the distribution of pixels, so we can say that 7.6 Image Transforms 469 127 50 0 -128 -50 255 255 128 128 0 0 (a) (b) Figure 7.17: Rotating a Cloud of Points. 470 7. Image Compression 0 50 100 150 200 250 300 0 10 20 30 40 50 60 70 80 90 100 0 50 100 150 200 250 300 0 10 20 30 40 50 60 70 80 90 100 0 50 100 150 200 250 300 350 0 10 20 30 40 50 60 70 80 90 0 50 100 150 200 250 300 0 100 200 300 400 500 600 700 800 900 1000 (a) (b) (d) (c) filename=’lena128’; dim=128; xdist=zeros(256,1); ydist=zeros(256,1); fid=fopen(filename,’r’); img=fread(fid,[dim,dim])’; for col=1:2:dim-1 for row=1:dim x=img(row,col)+1; y=img(row,col+1)+1; xdist(x)=xdist(x)+1; ydist(y)=ydist(y)+1; end end figure(1), plot(xdist), colormap(gray) %dist of x&y values figure(2), plot(ydist), colormap(gray) %before rotation xdist=zeros(325,1); % clear arrays ydist=zeros(256,1); for col=1:2:dim-1 for row=1:dim x=round((img(row,col)+img(row,col+1))*0.7071); y=round((-img(row,col)+img(row,col+1))*0.7071)+101; xdist(x)=xdist(x)+1; ydist(y)=ydist(y)+1; end end figure(3), plot(xdist), colormap(gray) %dist of x&y values figure(4), plot(ydist), colormap(gray) %after rotation Figure 7.18: Distribution of Image Pixels Before and After Rotation. 7.6 Image Transforms 471 the rotation has concentrated (or compacted) the energy in the x coordinate and has created compression this way. Concentrating the energy in one coordinate has another advantage. It makes it possible to quantize that coordinate more finely than the other coordinates. This type of quantization results in better (lossy) compression. The following simple example illustrates the power of this basic transform. We start with the point (4, 5), whose two coordinates are similar. Using Equation (7.4) the point is transformed to (4, 5)R = (9, 1)/.2 . (6.36396, 0.7071). The energies of the point and its transform are 42 + 52 = 41 = (92 + 12)/2. If we delete the smaller coordinate (4) of the point, we end up with an error of 42/41 = 0.39. If, on the other hand, we delete the smaller of the two transform coefficients (0.7071), the resulting error is just 0.70712/41 = 0.012. Another way to obtain the same error is to consider the reconstructed point. Passing 1 .2 (9, 1) through the inverse transform [Equation (7.5)] results in the original point (4, 5). Doing the same with 1 .2 (9, 0) results in the approximate reconstructed point (4.5, 4.5). The energy difference between the original and reconstructed points is the same small quantity (42 + 52) - (4.52 + 4.52) 42 + 52  = 41 - 40.5 41 = 0.0012. This simple transform can easily be extended to any number of dimensions. Instead of selecting pairs of adjacent pixels we can select triplets. Each triplet becomes a point in three-dimensional space, and these points form a cloud concentrated around the line that forms equal (although not 45.) angles with the three coordinate axes. When this line is rotated such that it coincides with the x axis, the y and z coordinates of the transformed points become small numbers. The transformation is done by multiplying each point by a 3Χ3 rotation matrix, and such a matrix is, of course, orthonormal. The transformed points are then separated into three coefficient vectors, of which the last two consist of small numbers. For maximum compression each coefficient vector should be quantized separately. This can be extended to more than three dimensions, with the only difference being that we cannot visualize spaces of dimensions higher than three. However, the mathematics can easily be extended. Some compression methods, such as JPEG, divide an image into blocks of 8 Χ 8 pixels each, and rotate first each row then each column, by means of Equation (7.16), as shown in Section 7.8. This double rotation produces a set of 64 transformed values, of which the first—termed the “DC coefficient”—is large, and the other 63 (called the “AC coefficients”) are normally small. Thus, this transform concentrates the energy in the first of 64 dimensions. The set of DC coefficients and each of the sets of 63 AC coefficients should, in principle, be quantized separately (JPEG does this a little differently, though; see Section 7.10.3). 472 7. Image Compression 7.7 Orthogonal Transforms Image transforms are designed to have two properties: (1) reduce image redundancy by reducing the sizes of most pixels and (2) identify the less important parts of the image by isolating the various frequencies of the image. Thus, this section starts with a short discussion of frequencies. We intuitively associate a frequency with a wave. Water waves, sound waves, and electromagnetic waves have frequencies, but pixels in an image can also feature frequencies. Figure 7.19 shows a small, 5Χ8 bi-level image that illustrates this concept. The top row is uniform, so we can assign it zero frequency. The rows below it have increasing pixel frequencies as measured by the number of color changes along a row. The four waves on the right roughly correspond to the frequencies of the four top rows of the image. a b c d Figure 7.19: Image Frequencies. Image frequencies are important because of the following basic fact: Low frequencies correspond to the important image features, whereas high frequencies correspond to the details of the image, which are less important. Thus, when a transform isolates the various image frequencies, pixels that correspond to high frequencies can be quantized heavily, while pixels that correspond to low frequencies should be quantized lightly or not at all. This is how a transform can compress an image very effectively by losing information, but only information associated with unimportant image details. Practical image transforms should be fast and preferably also simple to implement. This suggests the use of linear transforms. In such a transform, each transformed value (or transform coefficient) ci is a weighted sum of the data items (the pixels) dj that are being transformed, where each item is multiplied by a weight wij . Thus, ci = j djwij , for i, j = 1, 2, . . . , n. For n = 4, this is expressed in matrix notation as follows: ... c1 c2 c3 c4 ... =... w11 w12 w13 w14 w21 w22 w23 w24 w31 w32 w33 w34 w41 w42 w43 w44 ... ... d1 d2 d3 d4 ... . For the general case, we can write C =W·D. Each row ofW is called a “basis vector.” The only quantities that have to be computed are the weights wij . The guiding principles are as follows: 1. Reducing redundancy. The first transform coefficient c1 can be large, but the remaining values c2, c3, . . . should be small. 2. Isolating frequencies. The first transform coefficient c1 should correspond to zero pixel frequency, and the remaining coefficients should correspond to higher and higher frequencies. 7.7 Orthogonal Transforms 473 The key to determining the weights wij is the fact that our data items dj are not arbitrary numbers but pixel values, which are nonnegative and correlated. The basic relation ci = j djwij suggests that the first coefficient c1 will be large if all the weights of the form w1j are positive. To make the other coefficients ci small, it is enough to make half the weights wij positive and the other half negative. A simple choice is to assign half the weights the value +1 and the other half the value -1. In the extreme case where all the pixels dj are identical, this will result in ci = 0. When the dj ’s are similar, ci will be small (positive or negative). This choice of wij satisfies the first requirement: to reduce pixel redundancy by means of a transform. In order to satisfy the second requirement, the weights wij of row i should feature frequencies that get higher with i. Weights w1j should have zero frequency; they should all be +1’s. Weights w2j should have one sign change; i.e., they should be +1, +1, . . . + 1,-1,-1, . . . ,-1. This continues until the last row of weights wnj should have the highest frequency +1,-1, +1,-1, . . . , +1,-1. The mathematical discipline of vector spaces coins the term “basis vectors” for our rows of weights. In addition to isolating the various frequencies of pixels dj , this choice results in basis vectors that are orthogonal. The basis vectors are the rows of matrix W, which is why this matrix and, by implication, the entire transform are also termed orthogonal. These considerations are satisfied by the orthogonal matrix W=... 1 1 1 1 1 1 -1 -1 1 -1 -1 1 1 -1 1 -1 ... . (7.6) The first basis vector (the top row of W) consists of all 1’s, so its frequency is zero. Each of the subsequent vectors has two +1’s and two -1’s, so they produce small transformed values, and their frequencies (measured as the number of sign changes along the basis vector) get higher. This matrix is similar to the Walsh-Hadamard transform [Equation (7.7)]. To illustrate how this matrix identifies the frequencies in a data vector, we multiply it by four vectors as follows: W·... 1001 ... =... 2020 ... , W·... 0 0.33 -0.33 -1 ... =... 0 2.66 0 1.33 ... , W·... 1000 ... =... 1111 ... , W·... 1 -0.8 1 -0.8 ... =... 0.4 00 3.6 ... . The results make sense when we discover how the four test vectors were determined (1, 0, 0, 1) = 0.5(1, 1, 1, 1) + 0.5(1,-1,-1, 1), (1, 0.33,-0.33,-1) = 0.66(1, 1,-1,-1) + 0.33(1,-1, 1,-1), (1, 0, 0, 0) = 0.25(1, 1, 1, 1) + 0.25(1, 1,-1,-1) + 0.25(1,-1,-1, 1) + 0.25(1,-1, 1,-1), (1,-0.8, 1,-0.8) = 0.1(1, 1, 1, 1) + 0.9(1,-1, 1,-1). The product of W and the first vector shows how that vector consists of equal amounts of the first and the third frequencies. Similarly, the transform (0.4, 0, 0, 3.6) shows that 474 7. Image Compression vector (1,-0.8, 1,-0.8) is a mixture of a small amount of the first frequency and nine times the fourth frequency. It is also possible to modify this transform to conserve the energy of the data vector. All that’s needed is to multiply the transformation matrix W by the scale factor 1/2. Thus, the product (W/2)Χ(a, b, c, d) has the same energy a2 + b2 + c2 + d2 as the data vector (a, b, c, d). An example is the product of W/2 and the correlated vector (5, 6, 7, 8). It results in the transform coefficients (13,-2, 0,-1), where the first coefficient is large and the remaining ones are smaller than the original data items. The energy of both (5, 6, 7, 8) and (13,-2, 0,-1) is 174, but whereas in the former vector the first component accounts for only 14% of the energy, in the transformed vector the first component accounts for 97% of the energy. This is how our simple orthogonal transform compacts the energy of the data vector. Another advantage ofWis that it also performs the inverse transform. The product (W/2)·(13,-2, 0,-1)T reconstructs the original data (5, 6, 7, 8). We are now in a position to appreciate the compression potential of this transform. We use matrix W/2 to transform the (not very correlated) data vector d = (4, 6, 5, 2). The result is t = (8.5, 1.5,-2.5, 0.5). It’s easy to transform t back to d, but t itself does not provide any compression. In order to achieve compression, we quantize the components of t, and the point is that even after heavy quantization, it is still possible to get back a vector very similar to the original d. We first quantize t to the integers (9, 1,-3, 0) and perform the inverse transform to get back (3.5, 6.5, 5.5, 2.5). In a similar experiment, we completely delete the two smallest elements and inverse-transform the coarsely quantized vector (8.5, 0,-2.5, 0). This produces the reconstructed data (3, 5.5, 5.5, 3), still very close to the original values of d. The conclusion is that even this simple, intuitive transform is a powerful tool for “squeezing out” the redundancy in data. More sophisticated transforms produce results that can be quantized coarsely and still be used to reconstruct the original data to a high degree. 7.7.1 Two-Dimensional Transforms Given two-dimensional data such as the 4Χ4 matrix D =... 5 6 7 4 6 5 7 5 7 7 6 6 8 8 8 8 ... , where each of the four columns is highly correlated, we can apply our simple onedimensional transform to the columns of D. The result is C =W·D =... 1 1 1 1 1 1 -1 -1 1 -1 -1 1 1 -1 1 -1 ... ·D =... 26 26 28 23 -4 -4 0 -5 0 2 2 1 -2 0 -2 -3 ... . Each column of C is the transform of a column of D. Notice how the top element of each column of C is dominant, because the data in the corresponding column of 7.7 Orthogonal Transforms 475 D is correlated. Notice also that the rows of C are still correlated. C is the first stage in a two-stage process that produces the two-dimensional transform of matrix D. The second stage should transform each row of C, and this is done by multiplying C by the transpose WT . Our particular W, however, is symmetric, so we end up with C = C ·WT =W·D·WT =W·D·W or C =..... 26 26 28 23 -4 -4 0 -5 0 2 2 1 -2 0 -2 -3 ..... ... 1 1 1 1 1 1 -1 -1 1 -1 -1 1 1 -1 1 -1 ... =... 103 1 -5 5 -13 -3 -5 5 5 -1 -3 -1 -7 3 -3 -1 ... . The elements of C are decorrelated. The top-left element is dominant. It contains most of the total energy of the original D. The elements in the top row and the leftmost column are somewhat large, while the remaining elements are smaller than the original data items. The double-stage, two-dimensional transformation has reduced the correlation in both the horizontal and vertical dimensions. As in the one-dimensional case, excellent compression can be achieved by quantizing the elements of C, especially those that correspond to higher frequencies (i.e., located toward the bottom-right corner of C). This is the essence of orthogonal transforms. The remainder of this section discusses the following important transforms: 1. The Walsh-Hadamard transform (WHT, Section 7.7.2) is fast and easy to compute (it requires only additions and subtractions), but its performance, in terms of energy compaction, is lower than that of the DCT. 2. The Haar transform [Stollnitz et al. 96] is a simple, fast transform. It is the simplest wavelet transform and is discussed in Section 7.7.3 and in Chapter 8. 3. The Karhunen-Lo`eve transform (KLT, Section 7.7.4) is the best one theoretically, in the sense of energy compaction (or, equivalently, pixel decorrelation). However, its coefficients are not fixed; they depend on the data to be compressed. Calculating these coefficients (the basis of the transform) is slow, as is the calculation of the transformed values themselves. Since the coefficients are data dependent, they have to be included in the compressed stream. For these reasons and because the DCT performs almost as well, the KLT is not generally used in practice. 4. The discrete cosine transform (DCT) is discussed in detail in Section 7.8. This important transform is almost as efficient as the KLT in terms of energy compaction, but it uses a fixed basis, independent of the data. There are also fast methods for calculating the DCT. This method is used by JPEG and MPEG audio. 7.7.2 Walsh-Hadamard Transform As mentioned earlier, this transform has low compression efficiency, which is why it is not used much in practice. It is, however, fast, because it can be computed with just additions, subtractions, and an occasional right shift (to replace a division by a power of 2). 476 7. Image Compression Given an NΧN block of pixels Pxy (where N must be a power of 2, N = 2n), its two-dimensional WHT and inverse WHT are defined by Equations (7.7) and (7.8): H(u, v) = N-1 x=0 N-1 y=0 pxyg(x, y, u, v) = 1 N N-1 x=0 N-1 y=0 pxy(-1)n-1 i=0 [bi(x)pi(u)+bi(y)pi(v)], (7.7) Pxy = N-1 u=0 N-1 v=0 H(u, v)h(x, y, u, v) = 1 N N-1 u=0 N-1 v=0 H(u, v)(-1)n-1 i=0 [bi(x)pi(u)+bi(y)pi(v)], (7.8) where H(u, v) are the results of the transform (i.e., the WHT coefficients), the quantity bi(u) is bit i of the binary representation of the integer u, and pi(u) is defined in terms of the bj(u) by Equation (7.9): p0(u) = bn-1(u), p1(u) = bn-1(u) + bn-2(u), p2(u) = bn-2(u) + bn-3(u), ... pn-1(u) = b1(u) + b0(u). (7.9) (Recall that n is defined above by N = 2n.) As an example, consider u = 6 = 1102. Bits zero, one, and two of 6 are 0, 1, and 1, respectively, so b0(6) = 0, b1(6) = 1, and b2(6) = 1. The quantities g(x, y, u, v) and h(x, y, u, v) are called the kernels (or basis images) of the WHT. These matrices are identical. Their elements are just +1 and -1, and they are multiplied by the factor 1 N . As a result, the WHT transform consists in multiplying each image pixel by +1 or -1, summing, and dividing the sum by N. Since N = 2n is a power of 2, dividing by it can be done by shifting n positions to the right. The WHT kernels are shown, in graphical form, for N = 4, in Figure 7.20, where white denotes +1 and black denotes -1 (the factor 1 N is ignored). The rows and columns of blocks in this figure correspond to values of u and v from 0 to 3, respectively. The rows and columns inside each block correspond to values of x and y from 0 to 3, respectively. The number of sign changes across a row or a column of a matrix is called the sequency of the row or column. The rows and columns in the figure are ordered in increased sequency. Some authors show similar but unordered figures, because this transform was defined by Walsh and by Hadamard in slightly different ways (see [Gonzalez and Woods 92] for more information). Compressing an image with the WHT is done similarly to the DCT, except that Equations (7.7) and (7.8) are used, instead of Equations (7.16) and (7.17). 7.7 Orthogonal Transforms 477 u v Figure 7.20: The Ordered WHT Kernel for N = 4.  Exercise 7.8: Use appropriate mathematical software to compute and display the basis images of the WHT for N = 8. 7.7.3 Haar Transform The Haar transform [Stollnitz et al. 92] is based on the Haar functions hk(x), which are defined for x . [0, 1] and for k = 0, 1, . . . , N - 1, where N = 2n. Its application is also discussed in Chapter 8. Before we discuss the actual transform, we have to mention that any integer k can be expressed as the sum k = 2p + q - 1, where 0 . p . n - 1, q = 0 or 1 for p = 0, and 1 . q . 2p for p = 0. For N = 4 = 22, for example, we get 0 = 20+0-1, 1 = 20+1-1, 2 = 21 + 1 - 1, and 3 = 21 + 2 - 1. The Haar basis functions are now defined as h0(x) def = h00(x) = 1 .N , for 0 . x . 1, (7.10) and hk(x) def = hpq(x) = 1 .N.. .. . 2p/2, q-1 2p . x < q-1/2 2p , -2p/2, q-1/2 2p . x < q 2p , 0, otherwise for x . [0, 1]. (7.11) 478 7. Image Compression The Haar transform matrix AN of order NΧN can now be constructed. A general element i, j of this matrix is the basis function hi(j), where i = 0, 1, . . . , N - 1 and j = 0/N, 1/N, . . . , (N - 1)/N. For example, A2 = h0(0/2) h0(1/2) h1(0/2) h1(1/2) = 1 .2 1 1 1 -1 (7.12) (recall that i = 1 implies p = 0 and q = 1). Figure 7.21 shows code to calculate this matrix for any N, and also the Haar basis images for N = 8.  Exercise 7.9: Compute the Haar coefficient matrices A4 and A8. Given an image block P of order NΧN where N = 2n, its Haar transform is the matrix product ANPAN (Section 8.6). 7.7.4 Karhunen-Lo`eve Transform The Karhunen-Lo`eve transform (also called the Hotelling transform) has the best efficiency in the sense of energy compaction, but for the reasons mentioned earlier, it has more theoretical than practical value. Given an image, we break it up into k blocks of n pixels each, where n is typically 64 but can have other values, and k depends on the image size. We consider the blocks vectors and denote them by b(i), for i = 1, 2, . . . , k. The average vector is b = (i b(i))/k. A new set of vectors v(i) = b(i) - b is defined, causing the average (v(i))/k to be zero. We denote the nΧn KLT transform matrix that we are seeking by A. The result of transforming a vector v(i) is the weight vector w(i) = Av(i). The average of the w(i) is also zero. We now construct a matrix V whose columns are the v(i) vectors and another matrixWwhose columns are the weight vectors w(i): V = 2v(1), v(2), . . . , v(k)3, W= 2w(1),w(2), . . . ,w(k)3. Matrices V and W have n rows and k columns each. From the definition of w(i), we get W= A·V. The n coefficient vectors c(j) of the Karhunen-Lo`eve transform are given by c(j) = 2w(1) j ,w(2) j , . . . , w(k) j 3, j= 1, 2, . . . , n. Thus, vector c(j) consists of the jth elements of all the weight vectors w(i), for i = 1, 2, . . . , k (c(j) is the jth coordinate of the w(i) vectors). We now examine the elements of the matrix product W·WT (this is an n Χ n matrix). A general element in row a and column b of this matrix is the sum of products: W·WT ab = k i=1 w(i) a w(i) b = k i=1 c(a) i c(b) i = c(a) • c(b), for a, b . [1, n]. (7.13) The fact that the average of each w(i) is zero implies that a general diagonal element (W·WT )jj of the product matrix is the variance (up to a factor k) of the jth element 7.7 Orthogonal Transforms 479 Needs["GraphicsImage‘"] (* Draws 2D Haar Coefficients *) n=8; h[k_,x_]:=Module[{p,q}, If[k==0, 1/Sqrt[n], (* h_0(x) *) p=0; While[2^p<=k ,p++]; p--; q=k-2^p+1; (* if k>0, calc. p, q *) If[(q-1)/(2^p)<=x && x<(q-.5)/(2^p),2^(p/2), If[(q-.5)/(2^p)<=x && x0, for f = 0, 1, . . . , n - 1. The input is a set of n data values pt (pixels, audio samples, or other data), and the output is a set of n DCT transform coefficients (or weights) Gf . The first coefficient G0 is called the DC coefficient, and the rest are referred to as the AC coefficients (these terms have been inherited from electrical engineering, where they stand for “direct current” and “alternating current”). Notice that the coefficients are real numbers even if the input data consists of integers. Similarly, the coefficients may be positive or negative even if the input data consists of nonnegative numbers only. This computation is straightforward but slow (Section 7.8.5 discusses faster versions). The decoder inputs the DCT coefficients in sets of n and uses the inverse DCT (IDCT) to reconstruct the original data values (also in groups of n). The IDCT in one dimension is given by pt = 42 n n-1 j=0 CjGj cos (2t + 1)j. 2n , for t = 0, 1, . . . , n - 1. (7.15) The important feature of the DCT, the feature that makes it so useful in data compression, is that it takes correlated input data and concentrates its energy in just the first few transform coefficients. If the input data consists of correlated quantities, then most of the n transform coefficients produced by the DCT are zeros or small numbers, and only a few are large (normally the first ones). We will see that the early coefficients contain the important (low-frequency) image information and the later coefficients contain the less-important (high-frequency) image information. Compressing data with the DCT is therefore done by quantizing the coefficients. The small ones are quantized coarsely (possibly all the way to zero), and the large ones can be quantized finely to the nearest integer. After quantization, the coefficients (or variable-length codes 482 7. Image Compression assigned to the coefficients) are written on the compressed stream. Decompression is done by performing the inverse DCT on the quantized coefficients. This results in data items that are not identical to the original ones but are not much different. In practical applications, the data to be compressed is partitioned into sets of n items each and each set is DCT-transformed and quantized individually. The value of n is critical. Small values of n such as 3, 4, or 6 result in many small sets of data items. Such a small set is transformed to a small set of coefficients where the energy of the original data is concentrated in a few coefficients, but there are only a few coefficients in such a set! Thus, there are not enough small coefficients to quantize. Large values of n result in a few large sets of data. The problem in such a case is that the individual data items of a large set are normally not correlated and therefore result in a set of transform coefficients where all the coefficients are large. Experience indicates that n = 8 is a good value, and most data compression methods that employ the DCT use this value of n. The following experiment illustrates the power of the DCT in one dimension. We start with the set of eight correlated data items p = (12, 10, 8, 10, 12, 10, 8, 11), apply the DCT in one dimension to them, and find that it results in the eight coefficients 28.6375, 0.571202, 0.46194, 1.757, 3.18198, -1.72956, 0.191342, -0.308709. These can be fed to the IDCT and transformed by it to precisely reconstruct the original data (except for small errors caused by limited machine precision). Our goal, however, is to compress the data by quantizing the coefficients. We first quantize them to 28.6, 0.6, 0.5, 1.8, 3.2,-1.8, 0.2,-0.3, and apply the IDCT to get back 12.0254, 10.0233, 7.96054, 9.93097, 12.0164, 9.99321, 7.94354, 10.9989. We then quantize the coefficients even more, to 28, 1, 1, 2, 3,-2, 0, 0, and apply the IDCT to get back 12.1883, 10.2315, 7.74931, 9.20863, 11.7876, 9.54549, 7.82865, 10.6557. Finally, we quantize the coefficients to 28, 0, 0, 2, 3,-2, 0, 0, and still get back from the IDCT the sequence 11.236, 9.62443, 7.66286, 9.57302, 12.3471, 10.0146, 8.05304, 10.6842, where the largest difference between an original value (12) and a reconstructed one (11.236) is 0.764 (or 6.4% of 12). The code that does all that is listed in Figure 7.22. It seems magical that the eight original data items can be reconstructed to such high precision from just four transform coefficients. The explanation, however, relies on the following arguments instead of on magic: (1) The IDCT is given all eight transform coefficients, so it knows the positions, not just the values, of the nonzero coefficients. (2) The first few coefficients (the large ones) contain the important information of the original data items. The small coefficients, the ones that are quantized heavily, contain less important information (in the case of images, they contain the image details). (3) The original data is redundant because of pixel correlation. 7.8 The Discrete Cosine Transform 483 n=8; p={12.,10.,8.,10.,12.,10.,8.,11.}; c=Table[If[t==1, 0.7071, 1], {t,1,n}]; dct[i_]:=Sqrt[2/n]c[[i+1]]Sum[p[[t+1]]Cos[(2t+1)i Pi/16],{t,0,n-1}]; q=Table[dct[i],{i,0,n-1}] (* use precise DCT coefficients *) q={28,0,0,2,3,-2,0,0}; (* or use quantized DCT coefficients *) idct[t_]:=Sqrt[2/n]Sum[c[[j+1]]q[[j+1]]Cos[(2t+1)j Pi/16],{j,0,n-1}]; ip=Table[idct[t],{t,0,n-1}] Figure 7.22: Experiments with the One-Dimensional DCT. The following experiment illustrates the performance of the DCT when applied to decorrelated data items. Given the eight decorrelated data items -12, 24, -181, 209, 57.8, 3, -184, and -250, their DCT produces -117.803, 166.823, -240.83, 126.887, 121.198, 9.02198, -109.496, -185.206. When these coefficients are quantized to (-120., 170.,-240., 125., 120., 9.,-110.,-185) and fed into the IDCT, the result is -12.1249, 25.4974, -179.852, 208.237, 55.5898, 0.364874, -185.42, -251.701, where the maximum difference (between 3 and 0.364874) is 2.63513 or 88% of 3. Obviously, even with such fine quantization the reconstruction is not as good as with correlated data.  Exercise 7.11: Compute the one-dimensional DCT [Equation (7.14)] of the eight correlated values 11, 22, 33, 44, 55, 66, 77, and 88. Show how to quantize them, and compute their IDCT from Equation (7.15). An important relative of the DCT is the Fourier transform (Section 8.1), which also has a discrete version termed the DFT. The DFT has important applications, but it does not perform well in data compression because it assumes that the data to be transformed is periodic. The following example illustrates the difference in performance between the DCT and the DFT. We start with the simple, highly correlated sequence of eight numbers (8, 16, 24, 32, 40, 48, 56, 64). It is displayed graphically in Figure 7.23a. Applying the DCT to it yields (100,-52, 0,-5, 0,-2, 0, 0.4). When this is quantized to (100,-52, 0,-5, 0, 0, 0, 0) and transformed back, it produces (8, 15, 24, 32, 40, 48, 57, 63), a sequence almost identical to the original input. Applying the DFT to the same input, on the other hand, yields (36, 10, 10, 6, 6, 4, 4, 4). When this is quantized to (36, 10, 10, 6, 0, 0, 0, 0) and is transformed back, it produces (24, 12, 20, 32, 40, 51, 59, 48). This output is shown in Figure 7.23b, and it illustrates the tendency of the Fourier transform to produce a periodic result. The DCT in one dimension can be used to compress one-dimensional data, such as audio samples. This chapter, however, discusses image compression which is based on the two-dimensional correlation of pixels (a pixel tends to resemble all its near neighbors, 484 7. Image Compression (a) (b) Figure 7.23: (a) One-Dimensional Input. (b) Its Inverse DFT. not just those in its row). This is why practical image compression methods use the DCT in two dimensions. This version of the DCT is applied to small parts (data blocks) of the image. It is computed by applying the DCT in one dimension to each row of a data block, then to each column of the result. Because of the special way the DCT in two dimensions is computed, we say that it is separable in the two dimensions. Because it is applied to blocks of an image, we term it a “blocked transform.” It is defined by Gij = 4 2 m42 n CiCj n-1 x=0 m-1 y=0 pxy cos (2y + 1)j. 2m cos (2x + 1)i. 2n , (7.16) for 0 . i . n-1 and 0 . j . m-1 and for Ci and Cj defined by Equation (7.14). The first coefficient G00 is again termed the “DC coefficient,” and the remaining coefficients are called the “AC coefficients.” The image is broken up into blocks of nΧm pixels pxy (with n = m = 8 typically), and Equation (7.16) is used to produce a block of nΧm DCT coefficients Gij for each block of pixels. The coefficients are then quantized, which results in lossy but highly efficient compression. The decoder reconstructs a block of quantized data values by computing the IDCT whose definition is pxy = 4 2 m42 n n-1 i=0 m-1 j=0 CiCjGij cos (2x + 1)i. 2n cos (2y + 1)j. 2m , (7.17) where Cf = 1 .2 , f = 0 1 , f>0, for 0 . x . n - 1 and 0 . y . m - 1. We now show one way to compress an entire image with the DCT in several steps as follows: 1. The image is divided into k blocks of 8Χ8 pixels each. The pixels are denoted by pxy. If the number of image rows (columns) is not divisible by 8, the bottom row (rightmost column) is duplicated as many times as needed. 2. The DCT in two dimensions [Equation (7.16)] is applied to each block Bi. The result is a block (we’ll call it a vector) W(i) of 64 transform coefficients w(i) j (where j = 0, 1, . . . , 63). The k vectors W(i) become the rows of matrix W W=..... w(1) 0 w(1) 1 . . . w(1) 63 w(2) 0 w(2) 1 . . . w(2) 63 ... ... w(k) 0 w(k) 1 . . . w(k) 63 ..... . 7.8 The Discrete Cosine Transform 485 3. The 64 columns of W are denoted by C(0), C(1), . . . , C(63). The k elements of C(j) are 2w(1) j ,w(2) j , . . . , w(k) j 3. The first coefficient vector C(0) consists of the k DC coefficients. 4. Each vector C(j) is quantized separately to produce a vector Q(j) of quantized coefficients (JPEG does this differently; see Section 7.10.3). The elements of Q(j) are then written on the compressed stream. In practice, variable-length codes are assigned to the elements, and the codes, rather than the elements themselves, are written on the compressed stream. Sometimes, as in the case of JPEG, variable-length codes are assigned to runs of zero coefficients, to achieve better compression. In practice, the DCT is used for lossy compression. For lossless compression (where the DCT coefficients are not quantized) the DCT is inefficient but can still be used, at least theoretically, because (1) most of the coefficients are small numbers and (2) there often are runs of zero coefficients. However, the small coefficients are real numbers, not integers, so it is not clear how to write them in full precision on the compressed stream and still have compression. Other image compression methods are better suited for lossless image compression. The decoder reads the 64 quantized coefficient vectors Q(j) of k elements each, saves them as the columns of a matrix, and considers the k rows of the matrix weight vectors W(i) of 64 elements each (notice that these W(i)’s are not identical to the original W(i)’s because of the quantization). It then applies the IDCT [Equation (7.17)] to each weight vector, to reconstruct (approximately) the 64 pixels of block Bi. (Again, JPEG does this differently.) We illustrate the performance of the DCT in two dimensions by applying it to two blocks of 8Χ8 values. The first block (Table 7.24a) has highly correlated integer values in the range [8, 12], and the second block has random values in the same range. The first block results in a large DC coefficient, followed by small AC coefficients (including 20 zeros, Table 7.24b, where negative numbers are underlined). When the coefficients are quantized (Table 7.24c), the result, shown in Table 7.24d, is very similar to the original values. In contrast, the coefficients for the second block (Table 7.25b) include just one zero. When quantized (Table 7.25c) and transformed back, many of the 64 results are very different from the original values (Table 7.25d).  Exercise 7.12: Show why the 64 values of Table 7.24 are correlated. The next example illustrates the difference in the performance of the DCT when applied to a continuous-tone image and to a discrete-tone image. We start with the highly correlated pattern of Table 7.26. This is an idealized example of a continuous-tone image, since adjacent pixels differ by a constant amount except the pixel (underlined) at row 7, column 7. The 64 DCT coefficients of this pattern are listed in Table 7.27. It is clear that there are only a few dominant coefficients. Table 7.28 lists the coefficients after they have been coarsely quantized, so that only four nonzero coefficients remain! The results of performing the IDCT on these quantized coefficients are shown in Table 7.29. It is obvious that the four nonzero coefficients have reconstructed the original pattern to a high degree. The only visible difference is in row 7, column 7, which has changed from 12 to 17.55 (marked in both figures). The Matlab code for this computation is listed in Figure 7.34. 486 7. Image Compression 12 10 8 10 12 10 8 11 11 12 10 8 10 12 10 8 8 11 12 10 8 10 12 10 10 8 11 12 10 8 10 12 12 10 8 11 12 10 8 10 10 12 10 8 11 12 10 8 8 10 12 10 8 11 12 10 10 8 10 12 10 8 11 12 81 0 0 0 0 0 0 0 0 1.57 0.61 1.90 0.38 1.81 0.20 0.32 0 0.61 0.71 0.35 0 0.07 0 0.02 0 1.90 0.35 4.76 0.77 3.39 0.25 0.54 0 0.38 0 0.77 8.00 0.51 0 0.07 0 1.81 0.07 3.39 0.51 1.57 0.56 0.25 0 0.20 0 0.25 0 0.56 0.71 0.29 0 0.32 0.02 0.54 0.07 0.25 0.29 0.90 (a) Original data (b) DCT coefficients 81 0 0 0 0 0 0 0 0 2 1 2 0 2 0 0 0 1 1 0 0 0 0 0 0 2 0 5 1 3 0 1 0 0 0 1 8 1 0 0 0 2 0 3 1 2 1 0 0 0 0 0 0 1 1 0 0 0 0 1 0 0 0 1 12.29 10.26 7.92 9.93 11.51 9.94 8.18 10.97 10.90 12.06 10.07 7.68 10.30 11.64 10.17 8.18 7.83 11.39 12.19 9.62 8.28 10.10 11.64 9.94 10.15 7.74 11.16 11.96 9.90 8.28 10.30 11.51 12.21 10.08 8.15 11.38 11.96 9.62 7.68 9.93 10.09 12.10 9.30 8.15 11.16 12.19 10.07 7.92 7.87 9.50 12.10 10.08 7.74 11.39 12.06 10.26 9.66 7.87 10.09 12.21 10.15 7.83 10.90 12.29 (c) Quantized (d) Reconstructed data (good) Table 7.24: Two-Dimensional DCT of a Block of Correlated Values. 8 10 9 11 11 9 9 12 11 8 12 8 11 10 11 10 9 11 9 10 12 9 9 8 9 12 10 8 8 9 8 9 12 8 9 9 12 10 8 11 8 11 10 12 9 12 12 10 10 10 12 10 12 10 10 12 12 9 11 11 9 8 8 12 79.12 0.98 0.64 1.51 0.62 0.86 1.22 0.32 0.15 1.64 0.09 1.23 0.10 3.29 1.08 2.97 1.26 0.29 3.27 1.69 0.51 1.13 1.52 1.33 1.27 0.25 0.67 0.15 1.63 1.94 0.47 1.30 2.12 0.67 0.07 0.79 0.13 1.40 0.16 0.15 2.68 1.08 1.99 1.93 1.77 0.35 0 0.80 1.20 2.10 0.98 0.87 1.55 0.59 0.98 2.76 2.24 0.55 0.29 0.75 2.40 0.05 0.06 1.14 (a) Original data (b) DCT coefficients 79 1 1 2 1 1 1 0 0 2 0 1 0 3 1 3 1 0 3 2 0 1 2 1 1 0 1 0 2 2 0 10 20 1 0 1 0 10 0 0 3 1 2 2 2 0 0 1 1 2 1 1 2 1 1 3 2 1 0 1 2 0 0 1 7.59 9.23 8.33 11.88 7.12 12.47 6.98 8.56 12.09 7.97 9.3 11.52 9.28 11.62 10.98 12.39 11.02 10.06 13.81 6.5 10.82 8.28 13.02 7.54 8.46 10.22 11.16 9.57 8.45 7.77 10.28 11.89 9.71 11.93 8.04 9.59 8.04 9.7 8.59 12.14 10.27 13.58 9.21 11.83 9.99 10.66 7.84 11.27 8.34 10.32 10.53 9.9 8.31 9.34 7.47 8.93 10.61 9.04 13.66 6.04 13.47 7.65 10.97 8.89 (c) Quantized (d) Reconstructed data (bad) Table 7.25: Two-Dimensional DCT of a Block of Random Values. Tables 7.30 through 7.33 show the same process applied to a Y-shaped pattern, typical of a discrete-tone image. The quantization, shown in Table 7.32, is light. The coefficients have only been truncated to the nearest integer. It is easy to see that the reconstruction, shown in Table 7.33, isn’t as good as before. Quantities that should have been 10 are between 8.96 and 10.11. Quantities that should have been zero are as big as 0.86. The conclusion is that the DCT performs well on continuous-tone images but is less efficient when applied to a discrete-tone image. 7.8.2 The DCT as a Basis The discussion so far has concentrated on how to use the DCT for compressing onedimensional and two-dimensional data. The aim of this section and the next one is to show why the DCT works the way it does and how Equations (7.14) and (7.16) have been derived. This topic is approached from two different directions. The first interpretation of the DCT is as a special basis of an n-dimensional vector space. We 7.8 The Discrete Cosine Transform 487 00 10 20 30 30 20 10 00 10 20 30 40 40 30 20 10 20 30 40 50 50 40 30 20 30 40 50 60 60 50 40 30 30 40 50 60 60 50 40 30 20 30 40 50 50 40 30 20 10 20 30 40 40 30 12 10 00 10 20 30 30 20 10 00 12 Table 7.26: A Continuous-Tone Pattern. 239 1.19 -89.76 -0.28 1.00 -1.39 -5.03 -0.79 1.18 -1.39 0.64 0.32 -1.18 1.63 -1.54 0.92 -89.76 0.64 -0.29 -0.15 0.54 -0.75 0.71 -0.43 -0.28 0.32 -0.15 -0.08 0.28 -0.38 0.36 -0.22 1.00 -1.18 0.54 0.28 -1.00 1.39 -1.31 0.79 -1.39 1.63 -0.75 -0.38 1.39 -1.92 1.81 -1.09 -5.03 -1.54 0.71 0.36 -1.31 1.81 -1.71 1.03 -0.79 0.92 -0.43 -0.22 0.79 -1.09 1.03 -0.62 Table 7.27: Its DCT Coefficients. 239 1 -90 0 0 0 0 0 0 0 0 0 0 0 0 0 -90 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 Table 7.28: Quantized Heavily to Just Four Nonzero Coefficients. 0.65 9.23 21.36 29.91 29.84 21.17 8.94 0.30 9.26 17.85 29.97 38.52 38.45 29.78 17.55 8.91 21.44 30.02 42.15 50.70 50.63 41.95 29.73 21.09 30.05 38.63 50.76 59.31 59.24 50.56 38.34 29.70 30.05 38.63 50.76 59.31 59.24 50.56 38.34 29.70 21.44 30.02 42.15 50.70 50.63 41.95 29.73 21.09 9.26 17.85 29.97 38.52 38.45 29.78 17.55 8.91 0.65 9.23 21.36 29.91 29.84 21.17 8.94 0.30 17 Table 7.29: Results of IDCT. 488 7. Image Compression 00 10 00 00 00 00 00 10 00 00 10 00 00 00 10 00 00 00 00 10 00 10 00 00 00 00 00 00 10 00 00 00 00 00 00 00 10 00 00 00 00 00 00 00 10 00 00 00 00 00 00 00 10 00 00 00 00 00 00 00 10 00 00 00 Table 7.30: A Discrete-Tone Image (Y). 13.75 -3.11 -8.17 2.46 3.75 -6.86 -3.38 6.59 4.19 -0.29 6.86 -6.85 -7.13 4.48 1.69 -7.28 1.63 0.19 6.40 -4.81 -2.99 -1.11 -0.88 -0.94 -0.61 0.54 5.12 -2.31 1.30 -6.04 -2.78 3.05 -1.25 0.52 2.99 -0.20 3.75 -7.39 -2.59 1.16 -0.41 0.18 0.65 1.03 3.87 -5.19 -0.71 -4.76 0.68 -0.15 -0.88 1.28 2.59 -1.92 1.10 -9.05 0.83 -0.21 -0.99 0.82 1.13 -0.08 1.31 -7.21 Table 7.31: Its DCT Coefficients. 13.75 -3 -8 2 3 -6 -3 6 4 -0 6 -6 -7 4 1 -7 1 0 6 -4 -2 -1 -0 -0 -0 0 5 -2 1 -6 -2 3 -1 0 2 -0 3 -7 -2 1 -0 0 0 1 3 -5 -0 -4 0 -0 -0 1 2 -1 1 -9 0 -0 -0 0 1 -0 1 -7 Table 7.32: Quantized Lightly by Truncating to Integer. -0.13 8.96 0.55 -0.27 0.27 0.86 0.15 9.22 0.32 0.22 9.10 0.40 0.84 -0.11 9.36 -0.14 0.00 0.62 -0.20 9.71 -1.30 8.57 0.28 -0.33 -0.58 0.44 0.78 0.71 10.11 1.14 0.44 -0.49 -0.39 0.67 0.07 0.38 8.82 0.09 0.28 0.41 0.34 0.11 0.26 0.18 8.93 0.41 0.47 0.37 0.09 -0.32 0.78 -0.20 9.78 0.05 -0.09 0.49 0.16 -0.83 0.09 0.12 9.15 -0.11 -0.08 0.01 Table 7.33: The IDCT. Bad Results. 7.8 The Discrete Cosine Transform 489 % 8x8 correlated values n=8; p=[00,10,20,30,30,20,10,00; 10,20,30,40,40,30,20,10; 20,30,40,50,50,40,30,20; ... 30,40,50,60,60,50,40,30; 30,40,50,60,60,50,40,30; 20,30,40,50,50,40,30,20; ... 10,20,30,40,40,30,12,10; 00,10,20,30,30,20,10,00]; figure(1), imagesc(p), colormap(gray), axis square, axis off dct=zeros(n,n); for j=0:7 for i=0:7 for x=0:7 for y=0:7 dct(i+1,j+1)=dct(i+1,j+1)+p(x+1,y+1)*cos((2*y+1)*j*pi/16)*cos((2*x+1)*i*pi/16); end; end; end; end; dct=dct/4; dct(1,:)=dct(1,:)*0.7071; dct(:,1)=dct(:,1)*0.7071; dct quant=[239,1,-90,0,0,0,0,0; 0,0,0,0,0,0,0,0; -90,0,0,0,0,0,0,0; 0,0,0,0,0,0,0,0; ... 0,0,0,0,0,0,0,0; 0,0,0,0,0,0,0,0; 0,0,0,0,0,0,0,0; 0,0,0,0,0,0,0,0]; idct=zeros(n,n); for x=0:7 for y=0:7 for i=0:7 if i==0 ci=0.7071; else ci=1; end; for j=0:7 if j==0 cj=0.7071; else cj=1; end; idct(x+1,y+1)=idct(x+1,y+1)+ ... ci*cj*quant(i+1,j+1)*cos((2*y+1)*j*pi/16)*cos((2*x+1)*i*pi/16); end; end; end; end; idct=idct/4; idct figure(2), imagesc(idct), colormap(gray), axis square, axis off Figure 7.34: Code for Highly Correlated Pattern. show that transforming a given data vector p by the DCT is equivalent to representing it by this special basis that isolates the various frequencies contained in the vector. Thus, the DCT coefficients resulting from the DCT transform of vector p indicate the various frequencies in the vector. The lower frequencies contain the important information in p, whereas the higher frequencies correspond to the details of the data in p and are therefore less important. This is why they can be quantized coarsely. The second interpretation of the DCT is as a rotation, as shown intuitively for n = 2 (two-dimensional points) in Figure 7.17. This interpretation considers the DCT a rotation matrix that rotates an n-dimensional point with identical coordinates (x,x, . . . , x) from its original location to the x axis, where its coordinates become (., 2, . . . , n) where the various i are small numbers or zeros. Both interpretations are illustrated for the case n = 3, because this is the largest number of dimensions where it is possible to visualize geometric transformations. 490 7. Image Compression For the special case n = 3, Equation (7.14) reduces to Gf = 42 3Cf 2 t=0 pt cos (2t + 1)f. 6 , for f = 0, 1, 2. Temporarily ignoring the normalization factors 2/3 and Cf , this can be written in matrix notation as .. G0 G1 G2 .. =.. cos 0 cos 0 cos 0 cos . 6 cos 3. 6 cos 5. 6 cos 2. 6 cos 23. 6 cos 25. 6 .. .. p0 p1 p2 .. = D· p. Thus, the DCT of the three data values p = (p0, p1, p2) is obtained as the product of the DCT matrix D and the vector p. We can therefore think of the DCT as the product of a DCT matrix and a data vector, where the matrix is constructed as follows: Select the three angles ./6, 3./6, and 5./6 and compute the three basis vectors cos(f.) for f = 0, 1, and 2, and for the three angles. The results are listed in Table 7.35 for the benefit of the reader. . 0.5236 1.5708 2.618 cos 0. 1. 1. 1. cos 1. 0.866 0 -0.866 cos 2. 0.5 -1 0.5 Table 7.35: The DCT Matrix for n = 3. Because of the particular choice of the three angles, these vectors are orthogonal but not orthonormal. Their magnitudes are .3, .1.5, and .1.5, respectively. Normalizing them results in the three vectors v1 = (0.5774, 0.5774, 0.5774), v2 = (0.7071, 0,-0.7071), and v3 = (0.4082,-0.8165, 0.4082). When stacked vertically, they produce the following 3Χ3 matrix M=.. 0.5774 0.5774 0.5774 0.7071 0 -0.7071 0.4082 -0.8165 0.4082.. . (7.18) [Equation (7.14) tells us how to normalize these vectors: Multiply each by 2/3, and then multiply the first by 1/.2.] Notice that as a result of the normalization the columns ofMhave also become orthonormal, soMis an orthonormal matrix (such matrices have special properties). The steps of computing the DCT matrix for an arbitrary n are as follows: 1. Select the n angles .j = (j+0.5)./n for j = 0, . . . , n-1. If we divide the interval [0, .] into n equal-size segments, these angles are the centerpoints of the segments. 2. Compute the n vectors vk for k = 0, 1, 2, . . . , n-1, each with the n components cos(k.j ). 3. Normalize each of the n vectors and arrange them as the n rows of a matrix. 7.8 The Discrete Cosine Transform 491 The angles selected for the DCT are .j = (j + 0.5)./n, so the components of each vector vk are cos[k(j + 0.5)./n] or cos[k(2j + 1)./(2n)]. Section 7.8.4 covers three other ways to select such angles. This choice of angles has the following two useful properties: (1) The resulting vectors are orthogonal, and (2) for increasing values of k, the n vectors vk contain increasing frequencies (Figure 7.36). For n = 3, the top row of M [Equation (7.18)] corresponds to zero frequency, the middle row (whose elements become monotonically smaller) represents low frequency, and the bottom row (with three elements that first go down, then up) represents high frequency. Given a threedimensional vector v = (v1, v2, v3), the product M·v is a triplet whose components indicate the magnitudes of the various frequencies included in v; they are frequency coefficients. [Strictly speaking, the product is M·vT , but we ignore the transpose in cases where the meaning is clear.] The following three extreme examples illustrate the meaning of this statement. 1 1 -0.5 -0.5 2 2 3 1.5 1.5 2.5 Figure 7.36: Increasing Frequencies. The first example is v = (v, v, v). The three components of v are identical, so they correspond to zero frequency. The product M·v produces the frequency coefficients (1.7322v, 0, 0), indicating no high frequencies. The second example is v = (v, 0,-v). The three components of v vary slowly from v to -v, so this vector contains a low frequency. The product M·v produces the coefficients (0, 1.4142v, 0), confirming this result. The third example is v = (v,-v, v). The three components of v vary from v to -v to v, so this vector contains a high frequency. The product M·v produces (0, 0, 1.6329v), again indicating the correct frequency. These examples are not very realistic because the vectors being tested are short, simple, and contain a single frequency each. Most vectors are more complex and contain several frequencies, which makes this method useful. A simple example of a vector with two frequencies is v = (1., 0.33,-0.34). The product M·v results in (0.572, 0.948, 0) which indicates a large medium frequency, small zero frequency, and no high frequency. This makes sense once we realize that the vector being tested is the sum 0.33(1, 1, 1) + 0.67(1, 0,-1). A similar example is the sum 0.9(-1, 1,-1)+0.1(1, 1, 1) = (-0.8, 1,-0.8), which when multiplied by M produces (-0.346, 0,-1.469). On the other hand, a vector with random components, such as (1, 0, 0.33), typically contains roughly equal amounts of all three frequencies and produces three large frequency coefficients. The product M·(1, 0, 0.33) produces (0.77, 0.47, 0.54) because (1, 0, 0.33) is the sum 0.33(1, 1, 1) + 0.33(1, 0,-1) + 0.33(1,-1, 1). 492 7. Image Compression Notice that if M·v = c, then MT ·c = M-1 ·c = v. The original vector v can therefore be reconstructed from its frequency coefficients (up to small differences due to the limited precision of machine arithmetic). The inverseM-1 ofMis also its transpose MT because M is orthonormal. A three-dimensional vector can have only three frequencies zero, medium, and high. Similarly, an n-dimensional vector can have n different frequencies, which this method can identify. We concentrate on the case n = 8 and start with the DCT in one dimension. Figure 7.37 shows eight cosine waves of the form cos(f.j ), for 0 . .j . ., with frequencies f = 0, 1, . . . , 7. Each wave is sampled at the eight points .j = . 16, 3. 16 , 5. 16 , 7. 16 , 9. 16 , 11. 16 , 13. 16 , 15. 16 (7.19) to form one basis vector vf , and the resulting eight vectors vf , f = 0, 1, . . . , 7 (a total of 64 numbers) are shown in Table 7.38 (Equation (9.6) is the normalized form). They serve as the basis matrix of the DCT. Notice the similarity between this table and matrix W of Equation (7.6). Because of the particular choice of the eight sample points, the vi’s are orthogonal. This is easy to check directly with appropriate mathematical software, but Section 7.8.4 describes a more elegant way of proving this property. After normalization, the vi’s can be considered either as an 8Χ8 transformation matrix (specifically, a rotation matrix, since it is orthonormal) or as a set of eight orthogonal vectors that constitute the basis of a vector space. Any vector p in this space can be expressed as a linear combination of the vi’s. As an example, we select the eight (correlated) numbers p = (0.6, 0.5, 0.4, 0.5, 0.6, 0.5, 0.4, 0.55) as our test data and express p as a linear combination p = wivi of the eight basis vectors vi. Solving this system of eight equations yields the eight weights w0 = 0.506, w1 = 0.0143, w2 = 0.0115, w3 = 0.0439, w4 = 0.0795, w5 = -0.0432, w6 = 0.00478, w7 = -0.0077. Weight w0 is not much different from the elements of p, but the other seven weights are much smaller. This is how the DCT (or any other orthogonal transform) can lead to compression. The eight weights can be quantized and written on the compressed stream, where they occupy less space than the eight components of p. Figure 7.39 illustrates this linear combination graphically. Each of the eight vi’s is shown as a row of eight small, gray rectangles (a basis image) where a value of +1 is painted white and -1 is black. The eight elements of vector p are also displayed as a row of eight grayscale pixels. To summarize, we interpret the DCT in one dimension as a set of basis images that have higher and higher frequencies. Given a data vector, the DCT separates the frequencies in the data and represents the vector as a linear combination (or a weighted sum) of the basis images. The weights are the DCT coefficients. This interpretation can be extended to the DCT in two dimensions. We apply Equation (7.16) to the case n = 8 to create 64 small basis images of 8 Χ 8 pixels each. The 64 images are then used as a basis of a 64-dimensional vector space. Any image B of 8Χ8 pixels can be expressed as 7.8 The Discrete Cosine Transform 493 1.5 2 2.5 3 3 3 0.5 1 0.5 1 1.5 2 2.5 0.5 1 1.5 2 2.5 0.5 1 1.5 2 2.5 3 0.5 1 1.5 2 2.5 3 0.5 1 1.5 2 2.5 3 0.5 1 1.5 2 2.5 3 0.5 1 1.5 2 2.5 3 -1 -0.5 0.5 1 -1 -0.5 0.5 1 -1 -0.5 0.5 1 -1 -0.5 0.5 1 -1 -0.5 0.5 1 -1 -0.5 0.5 1 -1 -0.5 0.5 1 1 0.5 1.5 2 Figure 7.37: Angle and Cosine Values for an 8-Point DCT. 494 7. Image Compression . 0.196 0.589 0.982 1.374 1.767 2.160 2.553 2.945 cos 0. 1. 1. 1. 1. 1. 1. 1. 1. cos 1. 0.981 0.831 0.556 0.195 -0.195 -0.556 -0.831 -0.981 cos 2. 0.924 0.383 -0.383 -0.924 -0.924 -0.383 0.383 0.924 cos 3. 0.831 -0.195 -0.981 -0.556 0.556 0.981 0.195 -0.831 cos 4. 0.707 -0.707 -0.707 0.707 0.707 -0.707 -0.707 0.707 cos 5. 0.556 -0.981 0.195 0.831 -0.831 -0.195 0.981 -0.556 cos 6. 0.383 -0.924 0.924 -0.383 -0.383 0.924 -0.924 0.383 cos 7. 0.195 -0.556 0.831 -0.981 0.981 -0.831 0.556 -0.195 Table 7.38: The Unnormalized DCT Matrix in One Dimension for n = 8. Table[N[t],{t,Pi/16,15Pi/16,Pi/8}] dctp[pw_]:=Table[N[Cos[pw t]],{t,Pi/16,15Pi/16,Pi/8}] dctp[0] dctp[1] ... dctp[7] Code for Table 7.38. dct[pw_]:=Plot[Cos[pw t], {t,0,Pi}, DisplayFunction->Identity, AspectRatio->Automatic]; dcdot[pw_]:=ListPlot[Table[{t,Cos[pw t]},{t,Pi/16,15Pi/16,Pi/8}], DisplayFunction->Identity] Show[dct[0],dcdot[0], Prolog->AbsolutePointSize[4], DisplayFunction->$DisplayFunction] ... Show[dct[7],dcdot[7], Prolog->AbsolutePointSize[4], DisplayFunction->$DisplayFunction] Code for Figure 7.37. 0.506 0.0143 0.0115 0.0439 0.0795 -0.0432 0.00478 -0.0077 p v0 v2 v4 v6 v7 wi Figure 7.39: A Graphic Representation of the One-Dimensional DCT. 7.8 The Discrete Cosine Transform 495 a linear combination of the basis images, and the 64 weights of this linear combination are the DCT coefficients of B. Figure 7.40 shows the graphic representation of the 64 basis images of the twodimensional DCT for n = 8. A general element (i, j) in this figure is the 8 Χ 8 image obtained by calculating the product cos(i · s) cos(j · t), where s and t are varied independently over the values listed in Equation (7.19) and i and j vary from 0 to 7. This figure can easily be generated by the Mathematica code shown with it. The alternative code shown is a modification of code in [Watson 94], and it requires the GraphicsImage.m package, which is not widely available. Using appropriate software, it is easy to perform DCT calculations and display the results graphically. Figure 7.41a shows a random 8Χ8 data unit consisting of zeros and ones. The same unit is shown in Figure 7.41b graphically, with 1 as white and 0 as black. Figure 7.41c shows the weights by which each of the 64 DCT basis images has to be multiplied in order to reproduce the original data unit. In this figure, zero is shown in neutral gray, positive numbers are bright (notice how bright the DC weight is), and negative numbers are shown as dark. Figure 7.41d shows the weights numerically. The Mathematica code that does all that is also listed. Figure 7.42 is similar, but for a very regular data unit.  Exercise 7.13: . Imagine an 8Χ8 block of values where all the odd-numbered rows consist of 1’s and all the even-numbered rows contain zeros. What can we say about the DCT weights of this block? 7.8.3 The DCT as a Rotation The second interpretation of matrix M [Equation (7.18)] is as a rotation. We already know that M·(v, v, v) results in (1.7322v, 0, 0) and this can be interpreted as a rotation of point (v, v, v) to the point (1.7322v, 0, 0). The former point is located on the line that makes equal angles with the three coordinate axes, and the latter point is on the x axis. When considered in terms of adjacent pixels, this rotation has a simple meaning. Imagine three adjacent pixels in an image. They are normally similar, so we start by examining the case where they are identical. When three identical pixels are considered the coordinates of a point in three dimensions, that point is located on the line x = y = z. Rotating this line to the x axis brings our point to that axis where its x coordinate hasn’t changed much and its y and z coordinates are zero. This is how such a rotation leads to compression. Generally, three adjacent pixels p1, p2, and p3 are similar but not identical, which locates the point (p1, p2, p3) somewhat off the line x = y = z. After the rotation, the point will end up near the x axis, where its y and z coordinates will be small numbers. This interpretation of M as a rotation makes sense because M is orthonormal and any orthonormal matrix is a rotation matrix. However, the determinant of a rotation matrix is 1, whereas the determinant of our matrix is -1. An orthonormal matrix whose determinant is -1 performs an improper rotation (a rotation combined with a reflection). To get a better insight into the transformation performed by M, we apply the QR matrix decomposition technique (Section 7.8.8) to decomposeMinto the matrix 496 7. Image Compression Figure 7.40: The 64 Basis Images of the Two-Dimensional DCT. dctp[fs_,ft_]:=Table[SetAccuracy[N[(1.-Cos[fs s]Cos[ft t])/2],3], {s,Pi/16,15Pi/16,Pi/8},{t,Pi/16,15Pi/16,Pi/8}]//TableForm dctp[0,0] dctp[0,1] ... dctp[7,7] Code for Figure 7.40. Needs["GraphicsImage‘"] (* Draws 2D DCT Coefficients *) DCTMatrix=Table[If[k==0,Sqrt[1/8],Sqrt[1/4]Cos[Pi(2j+1)k/16]], {k,0,7}, {j,0,7}] //N; DCTTensor=Array[Outer[Times, DCTMatrix[[#1]],DCTMatrix[[#2]]]&, {8,8}]; Show[GraphicsArray[Map[GraphicsImage[#, {-.25,.25}]&, DCTTensor,{2}]]] Alternative Code for Figure 7.40. 7.8 The Discrete Cosine Transform 497 10011101 11001011 01100100 00010010 01001011 11100110 11001011 01010010 (a) (b) (c) 4.000 -0.133 0.637 0.272 -0.250 -0.181 -1.076 0.026 0.081 -0.178 -0.300 0.230 0.694 -0.309 0.875 -0.127 0.462 0.125 0.095 0.291 0.868 -0.070 0.021 -0.280 0.837 -0.194 0.455 0.583 0.588 -0.281 0.448 0.383 -0.500 -0.635 -0.749 -0.346 0.750 0.557 -0.502 -0.540 -0.167 0 -0.366 0.146 0.393 0.448 0.577 -0.268 -0.191 0.648 -0.729 -0.008 -1.171 0.306 1.155 -0.744 0.122 -0.200 0.038 -0.118 0.138 -1.154 0.134 0.148 (d) Figure 7.41: An Example of the DCT in Two Dimensions. DCTMatrix=Table[If[k==0,Sqrt[1/8],Sqrt[1/4]Cos[Pi(2j+1)k/16]], {k,0,7}, {j,0,7}] //N; DCTTensor=Array[Outer[Times, DCTMatrix[[#1]],DCTMatrix[[#2]]]&, {8,8}]; img={{1,0,0,1,1,1,0,1},{1,1,0,0,1,0,1,1}, {0,1,1,0,0,1,0,0},{0,0,0,1,0,0,1,0}, {0,1,0,0,1,0,1,1},{1,1,1,0,0,1,1,0}, {1,1,0,0,1,0,1,1},{0,1,0,1,0,0,1,0}}; ShowImage[Reverse[img]] dctcoeff=Array[(Plus @@ Flatten[DCTTensor[[#1,#2]] img])&,{8,8}]; dctcoeff=SetAccuracy[dctcoeff,4]; dctcoeff=Chop[dctcoeff,.001]; MatrixForm[dctcoeff] ShowImage[Reverse[dctcoeff]] Code for Figure 7.41. 498 7. Image Compression 01010101 01010101 01010101 01010101 01010101 01010101 01010101 01010101 (a) (b) (c) 4.000 -0.721 0 -0.850 0 -1.273 0 -3.625 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 (d) Figure 7.42: An Example of the DCT in Two Dimensions. Some painters transform the sun into a yellow spot; others transform a yellow spot into the sun. —Pablo Picasso DCTMatrix=Table[If[k==0,Sqrt[1/8],Sqrt[1/4]Cos[Pi(2j+1)k/16]], {k,0,7}, {j,0,7}] //N; DCTTensor=Array[Outer[Times, DCTMatrix[[#1]],DCTMatrix[[#2]]]&, {8,8}]; img={{0,1,0,1,0,1,0,1},{0,1,0,1,0,1,0,1}, {0,1,0,1,0,1,0,1},{0,1,0,1,0,1,0,1},{0,1,0,1,0,1,0,1}, {0,1,0,1,0,1,0,1},{0,1,0,1,0,1,0,1},{0,1,0,1,0,1,0,1}}; ShowImage[Reverse[img]] dctcoeff=Array[(Plus @@ Flatten[DCTTensor[[#1,#2]] img])&,{8,8}]; dctcoeff=SetAccuracy[dctcoeff,4]; dctcoeff=Chop[dctcoeff,.001]; MatrixForm[dctcoeff] ShowImage[Reverse[dctcoeff]] Code for Figure 7.42. 7.8 The Discrete Cosine Transform 499 product T1ΧT2ΧT3ΧT4, where T1 =.. 0.8165 0 -0.5774 0 1 0 0.5774 0 0.8165.. , T2 =.. 0.7071 -0.7071 0 0.7071 0.7071 0 0 0 1 .. , T3 =.. 1 0 0 0 0 1 0 -1 0.. , T4 =.. 1 0 0 0 1 0 0 0 -1.. . Each of matrices T1, T2, and T3 performs a rotation about one of the coordinate axes (these are called Givens rotations). Matrix T4 is a reflection about the z axis. The transformation M·(1, 1, 1) can now be written as T1ΧT2ΧT3ΧT4Χ(1, 1, 1), where T4 reflects point (1, 1, 1) to (1, 1,-1), T3 rotates (1, 1,-1) 90. about the x axis to (1,-1,-1), which is rotated by T2 45. about the z axis to (1.4142, 0,-1), which is rotated by T1 35.26. about the y axis to (1.7321, 0, 0). [This particular sequence of transformations is a result of the order in which the individual steps of the QR decomposition have been performed. Performing the same steps in a different order results in different sequences of rotations. One example is (1) a reflection about the z axis that transforms (1, 1, 1) to (1, 1,-1), (2) a rotation of (1, 1,-1) 135. about the x axis to (1,-1.4142, 0), and (3) a further rotation of 54.74. about the z axis to (1.7321, 0, 0).] For an arbitrary n, this interpretation is similar. We start with a vector of n adjacent pixels. They are considered the coordinates of a point in n-dimensional space. If the pixels are similar, the point is located near the line that makes equal angles with all the coordinate axes. Applying the DCT in one dimension [Equation (7.14)] rotates the point and brings it close to the x axis, where its first coordinate hasn’t changed much and its remaining n - 1 coordinates are small numbers. This is how the DCT in one dimension can be considered a single rotation in n-dimensional space. The rotation can be broken up into a reflection followed by n - 1 Givens rotations, but a user of the DCT need not be concerned with these details. The DCT in two dimensions is interpreted similarly as a double rotation. This interpretation starts with a block of nΧn pixels (Figure 7.43a, where the pixels are labeled L). It first considers each row of this block as a point (px,0, px,1, . . . , px,n-1) in n-dimensional space, and it rotates the point by means of the innermost sum G1x,j = 42 n Cj n-1 y=0 pxy cos(2y + 1)j. 2n of Equation (7.16). This results in a block G1x,j of n Χ n coefficients where the first element of each row is dominant (labeled L in Figure 7.43b) and the remaining elements are small (labeled S in that figure). The outermost sum of Equation (7.16) is Gij = 42 n Ci n-1 x=0 G1x,j cos(2x + 1)i. 2n . 500 7. Image Compression Here, the columns of G1x,j are considered points in n-dimensional space and are rotated. The result is one large coefficient at the top-left corner of the block (L in Figure 7.43c) and n2 - 1 small coefficients elsewhere (S and s in that figure). This interpretation considers the two-dimensional DCT as two separate rotations in n dimensions; the first one rotates each of the n rows, and the second one rotates each of the n columns. It is interesting to note that 2n rotations in n dimensions are faster than one rotation in n2 dimensions, since the latter requires an n2Χn2 rotation matrix. L L L L L L L L L L L L L L L L L L L L L L L L L L L L L L L L L L L L L L L L L L L L L L L L L L L L L L L L L L L L L L L L L S S S S S S S L S S S S S S S L S S S S S S S L S S S S S S S L S S S S S S S L S S S S S S S L S S S S S S S L S S S S S S S L S S S S S S S S s s s s s s s S s s s s s s s S s s s s s s s S s s s s s s s S s s s s s s s S s s s s s s s S s s s s s s s (a) (b) (c) Figure 7.43: The Two-Dimensional DCT as a Double Rotation. 7.8.4 The Four DCT Types There are four ways to select n equally-spaced angles that generate orthogonal vectors of cosines. They correspond (after the vectors are normalized by scale factors) to four discrete cosine transforms designated DCT-1 through DCT-4. The most useful is DCT- 2, which is normally referred to as the DCT. Equation (7.20) lists the definitions of the four types. The actual angles of the DCT-1 and DCT-2 are listed (for n = 8) in Table 7.44. Note that DCT-3 is the transpose of DCT-2 and DCT-4 is a shifted version of DCT-1. Notice that the DCT-1 has n+1 vectors of n+1 cosines each. In each of the four types, the n (or n + 1) DCT vectors are orthogonal and become normalized after they are multiplied by the proper scale factor. Figure 7.45 lists Mathematica code to generate the normalized vectors of the four types and test for normalization. DCT1 k,j = 42 n CkCj cos k j . n , k,j= 0, 1, . . . , n, DCT2 k,j = 42 n Ck cos k(j + 1 2 ). n , k,j= 0, 1, . . . , n - 1, DCT3 k,j = 42 n Cj cos (k + 1 2 )j. n , k,j= 0, 1, . . . , n - 1, (7.20) DCT4 k,j = 42 n cos (k + 1 2 )(j + 1 2 ). n , k,j= 0, 1, . . . , n - 1, where the scale factor Cx is defined by Cx = 1/.2, if x = 0 or x = n, 1, otherwise. 7.8 The Discrete Cosine Transform 501 k scale DCT-1 Angles (9Χ9) 0 1 2.2 0* 0 0 0 0 0 0 0 0* 1 1 2 0 . 8 2. 8 3. 8 4. 8 5. 8 6. 8 7. 8 8. 8 2 1 2 0 2. 8 4. 8 6. 8 8. 8 10. 8 12. 8 14. 8 16. 8 3 1 2 0 3. 8 6. 8 9. 8 12. 8 15. 8 18. 8 21. 8 24. 8 4 1 2 0 4. 8 8. 8 12. 8 16. 8 20. 8 24. 8 28. 8 32. 8 5 1 2 0 5. 8 10. 8 15. 8 20. 8 25. 8 30. 8 35. 8 40. 8 6 1 2 0 6. 8 12. 8 18. 8 24. 8 30. 8 36. 8 42. 8 48. 8 7 1 2 0 7. 8 14. 8 21. 8 28. 8 35. 8 42. 8 49. 8 56. 8 8 1 2.2 0* 8. 8 16. 8 24. 8 32. 8 40. 8 48. 8 56. 8 64. 8 * * the scale factor for these four angles is 4. k scale DCT-2 Angles 0 1 2.2 0 0 0 0 0 0 0 0 1 1 2 . 16 3. 16 5. 16 7. 16 9. 16 11. 16 13. 16 15. 16 2 1 2 2. 16 6. 16 10. 16 14. 16 18. 16 22. 16 26. 16 30. 16 3 1 2 3. 16 9. 16 15. 16 21. 16 27. 16 33. 16 39. 16 45. 16 4 1 2 4. 16 12. 16 20. 16 28. 16 36. 16 44. 16 52. 16 60. 16 5 1 2 5. 16 15. 16 25. 16 35. 16 45. 16 55. 16 65. 16 75. 16 6 1 2 6. 16 18. 16 30. 16 42. 16 54. 16 66. 16 78. 16 90. 16 7 1 2 7. 16 21. 16 35. 16 49. 16 63. 16 77. 16 91. 16 105. 16 Table 7.44: Angle Values for the DCT-1 and DCT-2. Orthogonality can be proved either directly, by multiplying pairs of different vectors, or indirectly. The latter approach is discussed in detail in [Strang 99] and it proves that the DCT vectors are orthogonal by showing that they are the eigenvectors of certain symmetric matrices. In the case of the DCT-2, for example, the symmetric matrix is A =...... 1 -1 -1 2 -1 ... -1 2 -1 -1 1 ...... . For n = 3, matrix A3 =.. 1 -1 0 -1 2 -1 0 -1 1.. has eigenvectors (0.5774, 0.5774, 0.5774), (0.7071, 0,-0.7071), (0.4082,-0.8165, 0.4082) with eigenvalues 0, 1, and 3, respectively. Recall that these eigenvectors are the rows of matrix M of Equation (7.18). 502 7. Image Compression (* DCT-1. Notice (n+1)x(n+1) *) Clear[n, nor, kj, DCT1, T1]; n=8; nor=Sqrt[2/n]; kj[i_]:=If[i==0 || i==n, 1/Sqrt[2], 1]; DCT1[k_]:=Table[nor kj[j] kj[k] Cos[j k Pi/n], {j,0,n}] T1=Table[DCT1[k], {k,0,n}]; (* Compute nxn cosines *) MatrixForm[T1] (* display as a matrix *) (* multiply rows to show orthonormality *) MatrixForm[Table[Chop[N[T1[[i]].T1[[j]]]], {i,1,n}, {j,1,n}]] (* DCT-2 *) Clear[n, nor, kj, DCT2, T2]; n=8; nor=Sqrt[2/n]; kj[i_]:=If[i==0 || i==n, 1/Sqrt[2], 1]; DCT2[k_]:=Table[nor kj[k] Cos[(j+1/2)k Pi/n], {j,0,n-1}] T2=Table[DCT2[k], {k,0,n-1}]; (* Compute nxn cosines *) MatrixForm[T2] (* display as a matrix *) (* multiply rows to show orthonormality *) MatrixForm[Table[Chop[N[T2[[i]].T2[[j]]]], {i,1,n}, {j,1,n}]] (* DCT-3. This is the transpose of DCT-2 *) Clear[n, nor, kj, DCT3, T3]; n=8; nor=Sqrt[2/n]; kj[i_]:=If[i==0 || i==n, 1/Sqrt[2], 1]; DCT3[k_]:=Table[nor kj[j] Cos[(k+1/2)j Pi/n], {j,0,n-1}] T3=Table[DCT3[k], {k,0,n-1}]; (* Compute nxn cosines *) MatrixForm[T3] (* display as a matrix *) (* multiply rows to show orthonormality *) MatrixForm[Table[Chop[N[T3[[i]].T3[[j]]]], {i,1,n}, {j,1,n}]] (* DCT-4. This is DCT-1 shifted *) Clear[n, nor, DCT4, T4]; n=8; nor=Sqrt[2/n]; DCT4[k_]:=Table[nor Cos[(k+1/2)(j+1/2) Pi/n], {j,0,n-1}] T4=Table[DCT4[k], {k,0,n-1}]; (* Compute nxn cosines *) MatrixForm[T4] (* display as a matrix *) (* multiply rows to show orthonormality *) MatrixForm[Table[Chop[N[T4[[i]].T4[[j]]]], {i,1,n}, {j,1,n}]] Figure 7.45: Code for Four DCT Types. From the dictionary Exegete (EK-suh-jeet), noun: A person who explains or interprets difficult parts of written works. 7.8 The Discrete Cosine Transform 503 7.8.5 Practical DCT Equation (7.16) can be coded directly in any higher-level language. Since this equation is the basis of several compression methods such as JPEG and MPEG, its fast calculation is essential. It can be speeded up considerably by making several improvements, and this section offers some ideas. 1. Regardless of the image size, only 32 cosine functions are involved. They can be precomputed once and used as needed to calculate all the 8 Χ 8 data units. Calculating the expression pxy cos (2x + 1)i. 16 cos (2y + 1)j. 16  now amounts to performing two multiplications. Thus, the double sum of Equation (7.16) requires 64 Χ 2 = 128 multiplications and 63 additions.  Exercise 7.14: (Proposed by V. Saravanan.) Why are only 32 different cosine functions needed for the DCT? 2. A little algebraic tinkering shows that the double sum of Equation (7.16) can be written as the matrix product CPCT , where P is the 8 Χ 8 matrix of the pixels, C is the matrix defined by Cij = 5 1 .8 , i= 0 1 2 cos (2j+1)i. 16 , i>0, (7.21) and CT is the transpose of C. (The product of two matrices Amp and Bpn is a matrix Cmn defined by Cij = p k=1 aikbkj. For other properties of matrices, see any text on linear algebra.) Calculating one matrix element of the product CP therefore requires eight multiplications and seven (but for simplicity let’s say eight) additions. Multiplying the two 8Χ8 matrices C and P requires 64Χ8 = 83 multiplications and the same number of additions. Multiplying the product CP by CT requires the same number of operations, so the DCT of one 8Χ8 data unit requires 2Χ83 multiplications (and the same number of additions). Assuming that the entire image consists of nΧn pixels and that n = 8q, there are qΧq data units, so the DCT of all the data units requires 2q283 multiplications (and the same number of additions). In comparison, performing one DCT for the entire image would require 2n3 = 2q383 = (2q283)q operations. By dividing the image into data units, we reduce the number of multiplications (and also of additions) by a factor of q. Unfortunately, q cannot be too large, because that would mean very small data units. Recall that a color image consists of three components (often RGB, but sometimes YCbCr or YPbPr). In JPEG, the DCT is applied to each component separately, bringing the total number of arithmetic operations to 3Χ2q283 = 3,072q2. For a 512Χ512-pixel image, this implies 3072Χ642 = 12,582,912 multiplications (and the same number of additions). 504 7. Image Compression 3. Another way to speed up the DCT is to perform all the arithmetic operations on fixed-point (scaled integer) rather than on floating-point numbers. On many computers, operations on fixed-point numbers require (somewhat) sophisticated programming techniques, but they are considerably faster than floating-point operations (except on supercomputers, which are optimized for floating-point arithmetic). The DCT algorithm with smallest currently-known number of arithmetic operations is described in [Feig and Linzer 90]. Today, there are also various VLSI chips that perform this calculation efficiently. 7.8.6 The LLM Method This section describes the Loeffler-Ligtenberg-Moschytz (LLM) method for the DCT in one dimension [Loeffler et al. 89]. Developed in 1989 by Christoph Loeffler, Adriaan Ligtenberg, and George S. Moschytz, this algorithm computes the DCT in one dimension with a total of 29 additions and 11 multiplications. Recall that the DCT in one dimension involves multiplying a row vector by a matrix. For n = 8, multiplying the row by one column of the matrix requires eight multiplications and seven additions, so the total number of operations required for the entire operation is 64 multiplications and 56 additions. Reducing the number of multiplications from 64 to 11 represents a savings of 83% and reducing the number of additions from 56 to 29 represents a savings of 49%—very significant! Only the final result is listed here and the interested reader is referred to the original publication for the details. We start with the double sum of Equation (7.16) and claim that a little algebraic tinkering reduces it to the form CPCT , where P is the 8Χ8 matrix of the pixels, C is the matrix defined by Equation (7.21) and CT is the transpose of C. In the one-dimensional case, only one matrix multiplication, namely PC, is needed. The originators of this method show that matrix C can be written (up to a factor of .8) as the product of seven simple matrices, as shown in Figure 7.47. Even though the number of matrices has been increased, the problem has been simplified because our seven matrices are sparse and contain mostly 1’s and -1’s. Multiplying by 1 or by -1 does not require a multiplication, and multiplying something by 0 saves an addition. Table 7.46 summarizes the total number of arithmetic operations required to multiply a row vector by the seven matrices. Matrix Additions Multiplications C1 0 0 C2 8 12 C3 4 0 C4 2 0 C5 0 2 C6 4 0 C7 8 0 Total 26 14 Table 7.46: Number of Arithmetic Operations. 7.8 The Discrete Cosine Transform 505 C = ........... 1 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 1 0 0 1 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 1 ........... Χ ............. 1 1 0 0 0 0 0 0 1 -1 0 0 0 0 0 0 0 0 .2 cos 6. 16 .2 cos 2. 16 0 0 0 0 0 0 -.2 cos 2. 16 .2 cos 6. 16 0 0 0 0 0 0 0 0 .2 cos 7. 16 0 0 .2 cos . 16 0 0 0 0 0 .2 cos 3. 16 .2 cos 5. 16 0 0 0 0 0 0 -.2 cos 5. 16 .2 cos 3. 16 0 0 0 0 0 -.2 cos . 16 0 0 .2 cos 7. 16 ............. Χ ........... 1 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 1 -1 0 0 0 0 0 0 1 1 0 0 0 0 0 0 0 0 -1 1 0 0 0 0 0 0 1 1 ........... Χ ........... 1 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 1 -1 0 0 0 0 0 0 1 1 0 0 0 0 0 0 0 0 1 ........... Χ ........... 1 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 1/.2 0 0 0 0 0 0 0 0 1/.2 0 0 0 0 0 0 0 0 1 ........... Χ ........... 1 0 0 1 0 0 0 0 0 1 1 0 0 0 0 0 0 1 -1 0 0 0 0 0 1 0 0 -1 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 1 ........... Χ ........... 1 0 0 0 0 0 0 1 0 1 0 0 0 0 1 0 0 0 1 0 0 1 0 0 0 0 0 1 1 0 0 0 0 0 0 1 -1 0 0 0 0 0 1 0 0 -1 0 0 0 1 0 0 0 0 -1 0 1 0 0 0 0 0 0 -1 ........... = C1C2C3C4C5C6C7. Figure 7.47: Product of Seven Matrices. 506 7. Image Compression These surprisingly small numbers can be reduced further by the following observation. We notice that matrix C2 has three groups of four cosines each. One of the groups consists of (we ignore the .2) two cos 6 16. and two cos 2 16. (one with a negative sign). We use the trigonometric identity cos( . 2 - .) = sin . to replace the two ±cos 2 16. with ±sin 6 16.. Multiplying any matrix by C2 now results in products of the form Acos( 6 16.) - B sin( 6 16.) and B cos( 6 16.) + Asin( 6 16.). It seems that computing these two elements requires four multiplications and two additions (assuming that a subtraction takes the same time to execute as an addition). The following computation, however, yields the same result with three additions and three multiplications: T = (A + B) cos ., T - B(cos . - sin .), -T + A(cos . + sin .). Thus, the three groups now require nine additions and nine multiplications instead of the original 6 additions and 12 multiplications (two more additions are needed for the other nonzero elements of C2), which brings the totals of Table 7.46 down to 29 additions and 11 multiplications. There is no national science just as there is no national multiplication table; what is national is no longer science. —Anton Chekhov 7.8.7 Hardware Implementation of the DCT Table 7.44 lists the 64 angle values of the DCT-2 for n = 8. When the cosines of those angles are computed, we find that because of the symmetry of the cosine function, there are only six distinct nontrivial cosine values. They are summarized in Table 7.48, where a = 1/.2, bi = cos(i./16), and ci = cos(i./8). The six nontrivial values are b1, b3, b5, b7, c1, and c3. 1 1 1 1 1 1 1 1 b1 b3 b5 b7 -b7 -b5 -b3 -b1 c1 c3 -c3 -c1 -c1 -c3 c3 c1 b3 -b7 -b1 -b5 b5 b1 b7 -b3 a -a -a a a -a -a a b5 -b1 b7 b3 -b3 -b7 b1 -b5 c3 -c1 c1 -c3 -c3 c1 -c1 c3 b7 -b5 b3 -b1 b1 -b3 b5 -b7 Table 7.48: Six Distinct Cosine Values for the DCT-2. This feature can be exploited in a fast software implementation of the DCT or to make a simple hardware device to compute the DCT coefficients Gi for eight pixel values pi. Figure 7.49 shows how such a device may be organized in two parts, each computing four of the eight coefficients Gi. Part I is based on a 4Χ4 symmetric matrix whose elements are the four distinct bi’s. The eight pixels are divided into four groups of two pixels each. The two pixels of each group are subtracted, and the four differences become a row vector that’s multiplied by the four columns of the matrix to produce the 7.8 The Discrete Cosine Transform 507 four DCT coefficients G1, G3, G5, and G7. Part II is based on a similar 4Χ4 matrix whose nontrivial elements are the two ci’s. The computations are similar except that the two pixels of each group are added instead of subtracted. (p0 - p7), (p1 - p6), (p2 - p5), (p3 - p4)... b1 b3 b5 b7 b3 -b7 -b1 -b5 b5 -b1 b7 b3 b7 -b5 b3 -b1 ... › [G1,G3,G5,G7], (I) (p0 + p7), (p1 + p6), (p2 + p5), (p3 + p4)... 1 c1 a c3 1 c3 -a -c1 1 -c3 -a c1 1 -c1 a -c3 ... › [G0,G2,G4,G6]. (II) Figure 7.49: A Hardware Implementation of the DCT-2. Figure 7.50 (after [Chen et al. 77]) illustrates how such a device can be constructed out of simple adders, complementors, and multipliers. Notation such as c(3./16) in the figure refers to the cosine function.                                             .. .. .. .. .. .. .. ..  ..  ..  ..   ..   ..   ..  ..  ..  .. Figure 7.50: A Hardware Implementation of the DCT-2. 7.8.8 QR Matrix Decomposition This section provides background material on the technique of QR matrix decomposition. It is intended for those already familiar with matrices who want to master this method. Any matrix A can be factored into the matrix product QΧR, where Q is an orthogonal matrix and R is upper triangular. If A is also orthogonal, then R will also be orthogonal. However, an upper triangular matrix that’s also orthogonal must be diagonal. The orthogonality of R implies R-1 = RT and its being diagonal implies 508 7. Image Compression R-1ΧR = I. The conclusion is that if A is orthogonal, then R must satisfy RT ΧR = I, which means that its diagonal elements must be +1 or -1. If A = QΧR and R has this form, then A and Q are identical, except that columns i of A and Q will have opposite signs for all values of i where Ri,i = -1. The QR decomposition of matrix A into Q and R is done by a loop where each iteration converts one element of A to zero. When all the below-diagonal elements of A have been zeroed, it becomes the upper triangular matrix R. Each element Ai,j is zeroed by multiplying A by a Givens rotation matrix Ti,j . This is an antisymmetric matrix where the two diagonal elements Ti,i and Tj,j are set to the cosine of a certain angle ., and the two off-diagonal elements Tj,i and Ti,j are set to the sine and negative sine, respectively, of the same .. The sine and cosine of . are defined as cos . = Aj,j D , sin . = Ai,j D , where D = )A2 j,j + A2 i,j . Following are some examples of Givens rotation matrices:  c s -s c, .. 1 0 0 0 c s 0 -s c.. , ... 1 0 0 0 0 c 0 s 0 0 1 0 0 -s 0 c ... , ..... 1 0 0 0 0 0 c 0 s 0 0 0 1 0 0 0 -s 0 c 0 0 0 0 0 1 ..... . (7.22) Those familiar with rotation matrices will recognize that a Givens matrix [Givens 58] rotates a point through an angle whose sine and cosine are the s and c of Equation (7.22). In two dimensions, the rotation is done about the origin. In three dimensions, it is done about one of the coordinate axes [the x axis in Equation (7.22)]. In four dimensions, the rotation is about two of the four coordinate axes (the first and third in Equation (7.22)) and cannot be visualized. In general, an nΧn Givens matrix rotates a point about n-2 coordinate axes of an n-dimensional space. J. Wallace Givens, Jr. (1910–1993) pioneered the use of plane rotations in the early days of automatic matrix computations. Givens graduated from Lynchburg College in 1928, and he completed his Ph.D. at Princeton University in 1936. After spending three years at the Institute for Advanced Study in Princeton as an assistant of Oswald Veblen, Givens accepted an appointment at Cornell University, but later moved to Northwestern University. In addition to his academic career, Givens was the director of the Applied Mathematics Division at Argonne National Lab and, like his counterpart Alston Householder at Oak Ridge National Laboratory, Givens served as an early president of SIAM. He published his work on the rotations in 1958. —Carl D. Meyer Figure 7.51 is a Matlab function for the QR decomposition of a matrix A. Notice how Q is obtained as the product of the individual Givens matrices and how the double loop zeros all the below-diagonal elements column by column from the bottom up. 7.8 The Discrete Cosine Transform 509 function [Q,R]=QRdecompose(A); % Computes the QR decomposition of matrix A % R is an upper triangular matrix and Q % an orthogonal matrix such that A=Q*R. [m,n]=size(A); % determine the dimens of A Q=eye(m); % Q starts as the mxm identity matrix R=A; for p=1:n for q=(1+p):m w=sqrt(R(p,p)^2+R(q,p)^2); s=-R(q,p)/w; c=R(p,p)/w; U=eye(m); % Construct a U matrix for Givens rotation U(p,p)=c; U(q,p)=-s; U(p,q)=s; U(q,q)=c; R=U’*R; % one Givens rotation Q=Q*U; end end Figure 7.51: A Matlab Function for the QR Decomposition of a Matrix. “Computer!’ shouted Zaphod, “rotate angle of vision through oneeighty degrees and don’t talk about it!’ —Douglas Adams, The Hitchhikers Guide to the Galaxy 7.8.9 Vector Spaces The discrete cosine transform can also be interpreted as a change of basis in a vector space from the standard basis to the DCT basis, so this section is a short discussion of vector spaces, their relation to data compression and to the DCT, their bases, and the important operation of change of basis. An n-dimensional vector space is the set of all vectors of the form (v1, v2, . . . , vn). We limit the discussion to the case where the vi’s are real numbers. The attribute of vector spaces that makes them important to us is the existence of bases. Any vector (a, b, c) in three dimensions can be written as the linear combination (a, b, c) = a(1, 0, 0) + b(0, 1, 0) + c(0, 0, 1) = ai + bj + ck, so we say that the set of three vectors i, j, and k forms a basis of the three-dimensional vector space. Notice that the three basis vectors are orthogonal; the dot product of any two of them is zero. They are also orthonormal; the dot product of each with itself is 1. It is convenient to have an orthonormal basis, but this is not a requirement. The basis does not even have to be orthogonal. The set of three vectors i, j, and k can be extended to any number of dimensions. A basis for an n-dimensional vector space may consist of the n vectors vi for i = 1, 2, . . . , n, where element j of vector vi is the Kronecker delta function .ij . This simple basis is the standard basis of the n-dimensional vector space. In addition to this basis, the ndimensional vector space can have other bases. We illustrate two other bases for n = 8. 510 7. Image Compression God made the integers, all else is the work of man. —Leopold Kronecker The DCT (unnormalized) basis consists of the eight vectors (1, 1, 1, 1, 1, 1, 1, 1), (1, 1, 1, 1,-1,-1,-1,-1), (1, 1,-1,-1,-1,-1, 1, 1), (1,-1,-1,-1, 1, 1, 1,-1), (1,-1,-1, 1, 1,-1,-1, 1), (1,-1, 1, 1,-1,-1, 1,-1), (1,-1, 1,-1,-1, 1,-1, 1), (1,-1, 1 - 1, 1,-1, 1,-1). Notice how their elements correspond to higher and higher frequencies. The (unnormalized) Haar wavelet basis (Section 8.6) consists of the eight vectors (1, 1, 1, 1, 1, 1, 1, 1), (1, 1, 1, 1,-1,-1,-1,-1), (1, 1,-1,-1, 0, 0, 0, 0), (0, 0, 0, 0, 1, 1,-1,-1), (1,-1, 0, 0, 0, 0, 0, 0), (0, 0, 1,-1, 0, 0, 0, 0), (0, 0, 0, 0, 1,-1, 0, 0), (0, 0, 0, 0, 0, 0, 1,-1). To understand why these bases are useful for data compression, recall that our data vectors are images or parts of images. The pixels of an image are normally correlated, but the standard basis takes no advantage of this. The vector of all 1’s, on the other hand, is included in the above bases because this single vector is sufficient to express any uniform image. Thus, a group of identical pixels (v, v, . . . , v) can be represented as the single coefficient v times the vector of all 1’s. [The discrete sine transform of Section 7.8.11 is unsuitable for data compression mainly because it does not include this uniform vector.] Basis vector (1, 1, 1, 1,-1,-1,-1,-1) can represent the energy of a group of pixels that’s half dark and half bright. Thus, the group (v, v, . . . , v,-v,-v, . . . ,-v) of pixels is represented by the single coefficient v times this basis vector. Successive basis vectors represent higher-frequency images, up to vector (1,-1, 1,-1, 1,-1, 1,-1). This basis vector resembles a checkerboard and therefore isolates the high-frequency details of an image. Those details are normally the least important and can be heavily quantized or even zeroed to achieve better compression. The vector members of a basis don’t have to be orthogonal. In order for a set S of vectors to be a basis, it has to have the following two properties: (1) The vectors have to be linearly independent and (2) it should be possible to express any member of the vector space as a linear combination of the vectors of S. For example, the three vectors (1, 1, 1), (0, 1, 0), and (0, 0, 1) are not orthogonal but form a basis for the threedimensional vector space. (1) They are linearly independent because none of them can be expressed as a linear combination of the other two. (2) Any vector (a, b, c) can be expressed as the linear combination a(1, 1, 1) + (b - a)(0, 1, 0) + (c - a)(0, 0, 1). Once we realize that a vector space may have many bases, we start looking for good bases. A good basis for data compression is one where the inverse of the basis matrix is easy to compute and where the energy of a data vector becomes concentrated in a few coefficients. The bases discussed so far are simple, being based on zeros and ones. 7.8 The Discrete Cosine Transform 511 The orthogonal bases have the added advantage that the inverse of the basis matrix is simply its transpose. Being fast is not enough, because the fastest thing we could do is to stay with the original standard basis. The reason for changing a basis is to get compression. The DCT base has the added advantage that it concentrates the energy of a vector of correlated values in a few coefficients. Without this property, there would be no reason to change the coefficients of a vector from the standard basis to the DCT basis. After changing to the DCT basis, many coefficients can be quantized, sometimes even zeroed, with a loss of only the least-important image information. If we quantize the original pixel values in the standard basis, we also achieve compression, but we lose image information that may be important. Once a basis has been selected, it is easy to express any given vector in terms of the basis vectors. Assuming that the basis vectors are bi and given an arbitrary vector P = (p1, p2, . . . , pn), we write P as a linear combination P = c1b1+c2b2+· · ·+cnbn of the bi’s with unknown coefficients ci. Using matrix notation, this is written P = c · B, where c is a row vector of the coefficients and B is the matrix whose rows are the basis vectors. The unknown coefficients can be computed by c = P · B-1 and this is the reason why a good basis is one where the inverse of the basis matrix is easy to compute. A simple example is the coefficients of a vector under the standard basis. We have seen that vector (a, b, c) can be written as the linear combination a(1, 0, 0)+b(0, 1, 0)+ c(0, 0, 1). Thus, when the standard basis is used, the coefficients of a vector P are simply its original elements. If we now want to compress the vector by changing to the DCT basis, we need to compute the coefficients under the new basis. This is an example of the important operation of change of basis. Given two bases bi and vi and assuming that a given vector P can be expressed as cibi and also as wivi, the problem of change of basis is to express one set of coefficients in terms of the other. Since the vectors vi constitute a basis, any vector can be expressed as a linear combination of them. Specifically, any bj can be written bj = i tijvi for some numbers tij . We now construct a matrix T from the tij and observe that it satisfies biT = vi for i = 1, 2, . . . , n. Thus, T is a linear transformation that transforms basis bi to vi. The numbers tij are the elements of T in basis vi. For our vector P, we can now write (cibi)T = civi, which implies n j=1 wjvj =j wjbjT =j wji vitij =i 2j wjtij3vi. This shows that ci = j tijwj ; in other words, a basis is changed by means of a linear transformation T and the same transformation also relates the elements of a vector in the old and new bases. Once we switch to a new basis in a vector space, every vector has new coordinates and every transformation has a different matrix. A linear transformation T operates on an n-dimensional vector and produces an m-dimensional vector, Thus, T(v) is a vector u. If m = 1, the transformation produces a scalar. If m = n-1, the transformation is a projection. Linear transformations satisfy the two important properties T(u+v) = T(u) +T(v) and T(cv) = cT(v). In general, the linear transformation of a linear combination T(c1v1+c2v2+· · ·+cnvn) equals the 512 7. Image Compression linear combination of the individual transformations c1T(v1)+c2T(v2)+· · ·+cnT(vn). This implies that the zero vector is transformed to itself under any linear transformation. Examples of linear transformations are projection, reflection, rotation, and differentiating a polynomial. The derivative of c1 + c2x + c3x2 is c2 + 2c3x. This is a transformation from the basis (c1, c2, c3) in three-dimensional space to basis (c2, c3) in two-dimensional space. The transformation matrix satisfies (c1, c2, c3)T = (c2, 2c3), so it is given by T =.. 0 0 1 0 0 2.. . Examples of nonlinear transformations are translation, the length of a vector, and adding a constant vector v0. The latter is nonlinear because if T(v) = v + v0 and we double the size of v, then T(2v) = 2v + v0 is different from 2T(v) = 2(v + v0). Transforming a vector v to its length ||v|| is also nonlinear because T(-v) = -T(v). Translation is nonlinear because it transforms the zero vector to a nonzero vector. In general, a linear transformation is performed by multiplying the transformed vector v by the transformation matrix T. Thus, u = v ·T or T(v) = v ·T. Notice that we denote by T both the transformation and its matrix. In order to describe a transformation uniquely, it is enough to describe what it does to the vectors of a basis. To see why this is true, we observe the following. If for a given vector v1 we know what T(v1) is, then we know what T(av1) is for any a. Similarly, if for a given v2 we know what T(v2) is, then we know what T(bv1) is for any b and also what T(av1 + bv2) is. Thus, we know how T transforms any vector in the plane containing v1 and v2. This argument shows that if we know what T(vi) is for all the vectors vi of a basis, then we know how T transforms any vector in the vector space. Given a basis bi for a vector space, we consider the special transformation that affects the magnitude of each vector but not its direction. Thus, T(bi) = .ibi for some number .i. The basis bi is the eigenvector basis of transformation T. Since we know T for the entire basis, we also know it for any other vector. Any vector v in the vector space can be expressed as a linear combination v = i cibi. If we apply our transformation to both sides and use the linearity property, we end up with T(v) = v · T =i cibi · T. (7.23) In the special case where v is the basis vector b1, Equation (7.23) implies T(b1) = i cibi · T. On the other hand, T(b1) = .1b1. We therefore conclude that c1 = .1 and, in general, that the transformation matrix T is diagonal with .i in position i of its diagonal. In the eigenvector basis, the transformation matrix is diagonal, so this is the perfect basis. We would love to have it in compression, but it is data dependent. It is called the Karhunen-Lo`eve transform (KLT) and is described in Section 7.7.4. 7.8.10 Rotations in Three Dimensions For those exegetes who want the complete story, the following paragraphs show how a proper rotation matrix (with a determinant of +1) that rotates a point (v, v, v) to the x axis can be derived from the general rotation matrix in three dimensions. 7.8 The Discrete Cosine Transform 513 A general rotation in three dimensions is fully specified by (1) an axis u of rotation, (2) the angle . of rotation, and (3) the direction (clockwise or counterclockwise as viewed from the origin) of the rotation about u. Given a unit vector u = (ux, uy, uz), matrixM of Equation (7.24) performs a rotation of .. about u. The rotation appears clockwise to an observer looking from the origin in the direction of u. If P = (x, y, z) is an arbitrary point, its position after the rotation is given by the product P ·M. M= (7.24) ... u2 x + cos .(1 - u2 x) uxuy(1 - cos .) - uz sin. uxuz(1 - cos .) + uy sin . uxuy(1 - cos .) + uz sin. u2 y + cos .(1 - u2 y) uyuz(1 - cos .) - ux sin . uxuz(1 - cos .) - uy sin. uyuz(1 - cos .) + ux sin. u2z + cos .(1 - u2z ) ... . The general rotation of Equation (7.24) can now be applied to our problem, which is to rotate the vector D = (1, 1, 1) to the x axis. The rotation should be done about the vector u that’s perpendicular to both D and (1, 0, 0). This vector is computed by the cross-product u = DΧ(1, 0, 0) = (0, 1,-1). Normalizing it yields u = (0, .,-.), where . = 1/.2. The next step is to compute the angle . between D and the x axis. This is done by normalizing D and computing the dot product of it and the x axis (recall that the dot product of two unit vectors is the cosine of the angle between them). The normalizedD is (ί, ί, ί), where ί = 1/.3, and the dot product results in cos . = ί, which also produces sin . = -1 - ί2 = -2/3 = -ί/.. The reason for the negative sign is that a rotation from (1, 1, 1) to (1, 0, 0) about u appears counterclockwise to an observer looking from the origin in the direction of positive u. The rotation matrix of Equation (7.24) was derived for the opposite direction of rotation. Also, cos . = ί implies that . = 54.76.. This angle, not 45., is the angle made by vector D with each of the three coordinate axes. [As an aside, when the number of dimensions increases, the angle between vector (1, 1, . . . , 1) and any of the coordinate axes approaches 90..] Substituting u, sin ., and cos . in Equation (7.24) and using the relations .2+ί(1- .2) = (ί + 1)/2 and -.2(1 - ί) = (ί - 1)/2 yields the simple rotation matrix M=.. ί -ί -ί ί .2 + ί(1 - .2) -.2(1 - ί) ί -.2(1 - ί) .2 + ί(1 - .2).. =.. ί -ί -ί ί (ί + 1)/2 (ί - 1)/2 ί (ί - 1)/2 (ί + 1)/2.. ... 0.5774 -0.5774 -0.5774 0.5774 0.7886 -0.2115 0.5774 -0.2115 0.7886.. . It is now easy to see that a point on the line x = y = z, with coordinates (v, v, v) is rotated by M to (v, v, v)M = (1.7322v, 0, 0). Notice that the determinant of M equals +1, so M is a rotation matrix, in contrast to the matrix of Equation (7.18), which generates improper rotations. 7.8.11 Discrete Sine Transform Readers who made it to this point may raise the question of why the cosine function, and not the sine, is used in the transform? Is it possible to use the sine function in a 514 7. Image Compression similar way to the DCT to create a discrete sine transform? Is there a DST, and if not, why? This short section discusses the differences between the sine and cosine functions and shows why these differences lead to a very ineffective discrete sine transform. A function f(x) that satisfies f(x) = -f(-x) is called odd. Similarly, a function for which f(x) = f(-x) is called even. For an odd function, it is always true that f(0) = -f(-0) = -f(0), so f(0) must be 0. Most functions are neither odd nor even, but the trigonometric functions sine and cosine are important examples of odd and even functions, respectively. Figure 7.52 shows that even though the only difference between them is phase (i.e., the cosine is a shifted version of the sine), this difference is enough to reverse their parity. When the (odd) sine curve is shifted, it becomes the (even) cosine curve, which has the same shape. -2. 2. Sine 0 1 Cosine -. . -1 Figure 7.52: The Sine and Cosine as Odd and Even Functions, Respectively. To understand the difference between the DCT and the DST, we examine the onedimensional case. The DCT in one dimension, Equation (7.14), employs the function cos[(2t + 1)f./16] for f = 0, 1, . . . , 7. For the first term, where f = 0, this function becomes cos(0), which is 1. This term is the familiar and important DC coefficient, which is proportional to the average of the eight data values being transformed. The DST is similarly based on the function sin[(2t + 1)f./16], resulting in a zero first term [since sin(0) = 0]. The first term contributes nothing to the transform, so the DST does not have a DC coefficient. The disadvantage of this can be seen when we consider the example of eight identical data values being transformed by the DCT and by the DST. Identical values are, of course, perfectly correlated. When plotted, they become a horizontal line. Applying the DCT to these values produces just a DC coefficient: All the AC coefficients are zero. The DCT compacts all the energy of the data into the single DC coefficient whose value is identical to the values of the data items. The IDCT can reconstruct the eight values perfectly (except for minor changes resulting from limited machine precision). Applying the DST to the same eight values, on the other hand, results in seven AC coefficients whose sum is a wave function that passes through the eight data points but oscillates between the points. This behavior, illustrated by Figure 7.53, has three disadvantages, namely (1) the energy of the original data values is not compacted, (2) the seven coefficients are not decorrelated (since the data values are perfectly correlated), and (3) quantizing the seven coefficients may greatly reduce the quality of the reconstruction done by the inverse DST. 7.8 The Discrete Cosine Transform 515 DST coefficients DCT coefficients Figure 7.53: The DCT and DST of Eight Identical Data Values. Example: Applying the DST to the eight identical values 100 results in the eight coefficients (0, 256.3, 0, 90, 0, 60.1, 0, 51). Using these coefficients, the IDST can reconstruct the original values, but it is easy to see that the AC coefficients do not behave like those of the DCT. They are not getting smaller, and there are no runs of zeros among them. Applying the DST to the eight highly correlated values 11, 22, 33, 44, 55, 66, 77, and 88 results in the even worse set of coefficients (0, 126.9,-57.5, 44.5,-31.1, 29.8,-23.8, 25.2). There is no energy compaction at all. These arguments and examples, together with the fact (discussed in [Ahmed et al. 74] and [Rao and Yip 90]) that the DCT produces highly decorrelated coefficients, argue strongly for the use of the DCT as opposed to the DST in data compression.  Exercise 7.15: Use mathematical software to compute and display the 64 basis images of the DST in two dimensions for n = 8. We are the wisp of straw, the plaything of the winds. We think that we are making for a goal deliberately chosen; destiny drives us towards another. Mathematics, the exaggerated preoccupation of my youth, did me hardly any service; and animals, which I avoided as much as ever I could, are the consolation of my old age. Nevertheless, I bear no grudge against the sine and the cosine, which I continue to hold in high esteem. They cost me many a pallid hour at one time, but they always afforded me some first rate entertainment: they still do so, when my head lies tossing sleeplessly on its pillow. —J. Henri Fabre, The Life of the Fly 516 7. Image Compression N=8; m=[1:N]’*ones(1,N); n=m’; % can also use cos instead of sin %A=sqrt(2/N)*cos(pi*(2*(n-1)+1).*(m-1)/(2*N)); A=sqrt(2/N)*sin(pi*(2*(n-1)+1).*(m-1)/(2*N)); A(1,:)=sqrt(1/N); C=A’; for row=1:N for col=1:N B=C(:,row)*C(:,col).’; %tensor product subplot(N,N,(row-1)*N+col) imagesc(B) drawnow end end Figure 7.54: The 64 Basis Images of the DST in Two Dimensions. 7.9 Test Images 517 7.9 Test Images New data compression methods that are developed and implemented have to be tested. Testing different methods on the same data makes it possible to compare their performance both in compression efficiency and in speed. This is why there are standard collections of test data, such as the Calgary Corpus and the Canterbury Corpus (mentioned in the Preface), and the ITU-T set of eight training documents for fax compression (Section 5.7.1). The need for standard test data has also been felt in the field of image compression, and there currently exist collections of still images commonly used by researchers and implementors in this field. Three of the four images shown here, namely Lena, mandril, and peppers, are arguably the most well known of them. They are continuous-tone images, although Lena has some features of a discrete-tone image. Each image is accompanied by a detail, showing individual pixels. It is easy to see why the peppers image is continuous-tone. Adjacent pixels that differ much in color are fairly rare in this image. Most neighboring pixels are very similar. In contrast, the mandril image, even though natural, is a bad example of a continuous-tone image. The detail (showing part of the right eye and the area around it) shows that many pixels differ considerably from their immediate neighbors because of the animal’s facial hair in this area. This image compresses badly under any compression method. However, the nose area, with mostly blue and red, is continuous-tone. The Lena image is mostly pure continuous-tone, especially the wall and the bare skin areas. The hat is good continuoustone, whereas the hair and the plume on the hat are bad continuous-tone. The straight lines on the wall and the curved parts of the mirror are features of a discrete-tone image. Figure 7.55: Lena and Detail. The Lena image is widely used by the image processing community, in addition to being popular in image compression. Because of the interest in it, its origin and 518 7. Image Compression Figure 7.56: Mandril and Detail. Figure 7.57: JPEG Blocking Artifacts. Figure 7.58: Peppers and Detail. 7.9 Test Images 519 Figure 7.59: A Discrete-Tone Image. Figure 7.60: A Discrete-Tone Image (Detail). 520 7. Image Compression history have been researched and are well documented. This image is part of the Playboy centerfold for November, 1972. It features the Swedish playmate Lena Soderberg (n΄ee Sjooblom), and it was discovered, clipped, and scanned in the early 1970s by an unknown researcher at the University of Southern California for use as a test image for his image compression research. It has since become the most important, well-known, and commonly used image in the history of imaging and electronic communications. As a result, Lena is currently considered by many the First Lady of the Internet. Playboy, which normally prosecutes unauthorized users of its images, has found out about the unusual use of one of its copyrighted images, but decided to give its blessing to this particular “application.” Lena herself currently lives in Sweden. She was told of her “fame” in 1988, was surprised and amused by it, and was invited to attend the 50th Anniversary IS&T (the society for Imaging Science and Technology) conference in Boston in May 1997. At the conference she autographed her picture, posed for new pictures (available on the www), and gave a presentation (about herself, not compression). The three images are widely available for downloading on the Internet. Figure 7.59 shows a typical discrete-tone image, with a detail shown in Figure 7.60. Notice the straight lines and the text, where certain characters appear several times (a source of redundancy). This particular image has few colors, but in general, a discretetone image may have many colors. Lena, Illinois, is a community of approximately 2,900 people. Lena is considered to be a clean and safe community located centrally to larger cities that offer other interests when needed. Lena is 2-1/2 miles from Lake Le-Aqua-Na State Park. The park offers hiking trails, fishing, swimming beach, boats, cross country skiing, horse back riding trails, as well as picnic and camping areas. It is a beautiful well-kept park that has free admission to the public. A great place for sledding and ice skating in the winter! (From http://www.villageoflena.com/) 7.10 JPEG JPEG is a sophisticated lossy/lossless compression method for color or grayscale still images (not videos). It does not handle bi-level (black and white) images very well. It also works best on continuous-tone images, where adjacent pixels have similar colors. An important feature of JPEG is its use of many parameters, allowing the user to adjust the amount of the data lost (and thus also the compression ratio) over a very wide range. Often, the eye cannot see any image degradation even at compression factors of 10 or 20. There are two operating modes, lossy (also called baseline) and lossless (which typically produces compression ratios of around 0.5). Most implementations support just the lossy mode. This mode includes progressive and hierarchical coding. A few of the many references to JPEG are [Pennebaker and Mitchell 92], [Wallace 91], and [Zhang 90]. JPEG is a compression method, not a complete standard for image representation. This is why it does not specify image features such as pixel aspect ratio, color space, or interleaving of bitmap rows. 7.10 JPEG 521 JPEG has been designed as a compression method for continuous-tone images. The main goals of JPEG compression are the following: 1. High compression ratios, especially in cases where image quality is judged as very good to excellent. 2. The use of many parameters, allowing knowledgeable users to experiment and achieve the desired compression/quality trade-off. 3. Obtaining good results with any kind of continuous-tone image, regardless of image dimensions, color spaces, pixel aspect ratios, or other image features. 4. A sophisticated, but not too complex compression method, allowing software and hardware implementations on many platforms. 5. Several modes of operation: (a) A sequential mode where each image component (color) is compressed in a single left-to-right, top-to-bottom scan; (b) A progressive mode where the image is compressed in multiple blocks (known as “scans”) to be viewed from coarse to fine detail; (c) A lossless mode that is important in cases where the user decides that no pixels should be lost (the trade-off is low compression ratio compared to the lossy modes); and (d) A hierarchical mode where the image is compressed at multiple resolutions allowing lower-resolution blocks to be viewed without first having to decompress the following higher-resolution blocks. The name JPEG is an acronym that stands for Joint Photographic Experts Group. This was a joint effort by the CCITT and the ISO (the International Standards Organization) that started in June 1987 and produced the first JPEG draft proposal in 1991. The JPEG standard has proved successful and has become widely used for image compression, especially in Web pages. The main JPEG compression steps are outlined here, and each step is then described in detail later. 1. Color images are transformed from RGB into a luminance/chrominance color space (Section 7.10.1; this step is skipped for grayscale images). The eye is sensitive to small changes in luminance but not in chrominance, so the chrominance part can later lose much data, and thus be highly compressed, without visually impairing the overall image quality much. This step is optional but important because the remainder of the algorithm works on each color component separately. Without transforming the color space, none of the three color components will tolerate much loss, leading to worse compression. 2. Color images are downsampled by creating low-resolution pixels from the original ones (this step is used only when hierarchical compression is selected; it is always skipped for grayscale images). The downsampling is not done for the luminance component. Downsampling is done either at a ratio of 2:1 both horizontally and vertically (the socalled 2h2v or 4:1:1 sampling) or at ratios of 2:1 horizontally and 1:1 vertically (2h1v or 4:2:2 sampling). Since this is done on two of the three color components, 2h2v reduces the image to 1/3 + (2/3) Χ (1/4) = 1/2 its original size, while 2h1v reduces it to 1/3 + (2/3) Χ (1/2) = 2/3 its original size. Since the luminance component is not touched, there is no noticeable loss of image quality. Grayscale images don’t go through this step. 3. The pixels of each color component are organized in groups of 8Χ8 pixels called data units, and each data unit is compressed separately. If the number of image rows or columns is not a multiple of 8, the bottom row and the rightmost column are duplicated 522 7. Image Compression as many times as necessary. In the noninterleaved mode, the encoder handles all the data units of the first image component, then the data units of the second component, and finally those of the third component. In the interleaved mode the encoder processes the three top-left data units of the three image components, then the three data units to their right, and so on. The fact that each data unit is compressed separately is one of the downsides of JPEG. If the user asks for maximum compression, the decompressed image may exhibit blocking artifacts due to differences between blocks. Figure 7.57 is an extreme example of this effect. 4. The discrete cosine transform (DCT, Section 7.8) is then applied to each data unit to create an 8Χ8 map of frequency components (Section 7.10.2). They represent the average pixel value and successive higher-frequency changes within the group. This prepares the image data for the crucial step of losing information. Since DCT involves the transcendental function cosine, it must involve some loss of information due to the limited precision of computer arithmetic. This means that even without the main lossy step (step 5 below), there will be some loss of image quality, but it is normally small. 5. Each of the 64 frequency components in a data unit is divided by a separate number called its quantization coefficient (QC), and then rounded to an integer (Section 7.10.3). This is where information is irretrievably lost. Large QCs cause more loss, so the highfrequency components typically have larger QCs. Each of the 64 QCs is a JPEG parameter and can, in principle, be specified by the user. In practice, most JPEG implementations use the QC tables recommended by the JPEG standard for the luminance and chrominance image components (Table 7.63). 6. The 64 quantized frequency coefficients (which are now integers) of each data unit are encoded using a combination of RLE and Huffman coding (Section 7.10.4). An arithmetic coding variant known as the QM coder (Section 5.11) can optionally be used instead of Huffman coding. 7. The last step adds headers and all the required JPEG parameters, and outputs the result. The compressed file may be in one of three formats (1) the interchange format, in which the file contains the compressed image and all the tables needed by the decoder (mostly quantization tables and tables of Huffman codes), (2) the abbreviated format for compressed image data, where the file contains the compressed image and may contain no tables (or just a few tables), and (3) the abbreviated format for table-specification data, where the file contains just tables, and no compressed image. The second format makes sense in cases where the same encoder/decoder pair is used, and they have the same tables built in. The third format is used in cases where many images have been compressed by the same encoder, using the same tables. When those images need to be decompressed, they are sent to a decoder preceded by one file with table-specification data. The JPEG decoder performs the reverse steps. (Thus, JPEG is a symmetric compression method.) The progressive mode is a JPEG option. In this mode, higher-frequency DCT coefficients are written on the compressed stream in blocks called “scans.” Each scan that is read and processed by the decoder results in a sharper image. The idea is to use the first few scans to quickly create a low-quality, blurred preview of the image, and then either input the remaining scans or stop the process and reject the image. The trade-off is that the encoder has to save all the coefficients of all the data units in a 7.10 JPEG 523 memory buffer before they are sent in scans, and also go through all the steps for each scan, slowing down the progressive mode. Figure 7.61a shows an example of an image with resolution 1024Χ512. The image is divided into 128Χ64 = 8192 data units, and each is transformed by the DCT, becoming a set of 64 8-bit numbers. Figure 7.61b is a block whose depth corresponds to the 8,192 data units, whose height corresponds to the 64 DCT coefficients (the DC coefficient is the top one, numbered 0), and whose width corresponds to the eight bits of each coefficient. After preparing all the data units in a memory buffer, the encoder writes them on the compressed stream in one of two methods, spectral selection or successive approximation (Figure 7.61c,d). The first scan in either method is the set of DC coefficients. If spectral selection is used, each successive scan consists of several consecutive (a band of) AC coefficients. If successive approximation is used, the second scan consists of the four most-significant bits of all AC coefficients, and each of the following four scans, numbers 3 through 6, adds one more significant bit (bits 3 through 0, respectively). In the hierarchical mode, the encoder stores the image several times in the output stream, at several resolutions. However, each high-resolution part uses information from the low-resolution parts of the output stream, so the total amount of information is less than that required to store the different resolutions separately. Each hierarchical part may use the progressive mode. The hierarchical mode is useful in cases where a high-resolution image needs to be output in low resolution. Older dot-matrix printers may be a good example of a low-resolution output device still in use. The lossless mode of JPEG (Section 7.10.5) calculates a “predicted” value for each pixel, generates the difference between the pixel and its predicted value (see Section 1.3.1 for relative encoding), and encodes the difference using the same method (i.e., Huffman or arithmetic coding) employed by step 5 above. The predicted value is calculated using values of pixels above and to the left of the current pixel (pixels that have already been input and encoded). The following sections discuss the steps in more detail: 7.10.1 Luminance The main international organization devoted to light and color is the International Committee on Illumination (Commission Internationale de l’΄Eclairage), abbreviated CIE. It is responsible for developing standards and definitions in this area. One of the early achievements of the CIE was its chromaticity diagram [Salomon 99], developed in 1931. It shows that no fewer than three parameters are required to define color. Expressing a certain color by the triplet (x, y, z) is similar to denoting a point in three-dimensional space, hence the term color space. The most common color space is RGB, where the three parameters are the intensities of red, green, and blue in a color. When used in computers, these parameters are normally in the range 0–255 (8 bits). 524 7. Image Compression (c) 63 62 63 62 0 0 0 7 6 7 3 0 6 5 4 0 1 01 1 127 128 129 130 255 256 8192 8192 8065 8191 1 2 3 4 data units 1 2 012 2 (d) (a) (b) 63 62 012 63 62 012 63 62 012 7 6 1 0 1024=8Χ128 512=8Χ64 524,288 pixels 1st scan 1st scan 2nd scan 3rd scan 3rd scan 2nd scan 6th scan kth scan 8,192 data units Figure 7.61: Scans in the JPEG Progressive Mode. 7.10 JPEG 525 The CIE defines color as the perceptual result of light in the visible region of the spectrum, having wavelengths in the region of 400 nm to 700 nm, incident upon the retina (a nanometer, nm, equals 10-9 meter). Physical power (or radiance) is expressed in a spectral power distribution (SPD), often in 31 components each representing a 10-nm band. The CIE defines brightness as the attribute of a visual sensation according to which an area appears to emit more or less light. The brain’s perception of brightness is impossible to define, so the CIE defines a more practical quantity called luminance. It is defined as radiant power weighted by a spectral sensitivity function that is characteristic of vision (the eye is very sensitive to green, slightly less sensitive to red, and much less sensitive to blue). The luminous efficiency of the Standard Observer is defined by the CIE as a positive function of the wavelength, which has a maximum at about 555 nm. When a spectral power distribution is integrated using this function as a weighting function, the result is CIE luminance, which is denoted by Y. Luminance is an important quantity in the fields of digital image processing and compression. Luminance is proportional to the power of the light source. It is similar to intensity, but the spectral composition of luminance is related to the brightness sensitivity of human vision. The eye is very sensitive to small changes in luminance, which is why it is useful to have color spaces that use Y as one of their three parameters. A simple way to do this is to subtract Y from the Blue and Red components of RGB, and use the three components Y, B - Y, and R - Y as a new color space. The last two components are called chroma. They represent color in terms of the presence or absence of blue (Cb) and red (Cr) for a given luminance intensity. Various number ranges are used in B - Y and R - Y for different applications. The YPbPr ranges are optimized for component analog video. The YCbCr ranges are appropriate for component digital video such as studio video, JPEG, JPEG 2000, and MPEG. The YCbCr color space was developed as part of Recommendation ITU-R BT.601 (formerly CCIR 601) during the development of a worldwide digital component video standard. Y is defined to have a range of 16 to 235; Cb and Cr are defined to have a range of 16 to 240, with 128 equal to zero. There are several YCbCr sampling formats, such as 4:4:4, 4:2:2, 4:1:1, and 4:2:0, which are also described in the recommendation (see also Section 9.11.4). Conversions between RGB with a 16–235 range and YCbCr are linear and thus simple. Transforming RGB to YCbCr is done by (note the small weight of blue): Y = (77/256)R + (150/256)G + (29/256)B, Cb = -(44/256)R - (87/256)G + (131/256)B + 128, Cr = (131/256)R - (110/256)G - (21/256)B + 128; while the opposite transformation is R = Y+1.371(Cr - 128), G = Y- 0.698(Cr - 128) - 0.336(Cb - 128), 526 7. Image Compression B = Y+1.732(Cb - 128). When performing YCbCr to RGB conversion, the resulting RGB values have a nominal range of 16–235, with possible occasional values in 0–15 and 236–255. Color and Human Vision An object has intrinsic and extrinsic attributes. They are defined as follows: An intrinsic attribute is innate, inherent, inseparable from the thing itself and independent of the way the rest of the world is. Examples of intrinsic attributes are length, shape, mass, electrical conductivity, and rigidity. Extrinsic is any attribute that is not contained in or belonging to an object. Examples are the monetary value, esthetic value, and usefulness of an object to us. In complete darkness, we cannot see. With light around us, we see objects and they have colors. Is color an intrinsic or an extrinsic attribute of an object? The surprising answer is neither. Stated another way, colors do not exist in nature. What does exist is light of different wavelengths. When such light enters our eye, the light-sensitive cells in the retina send signals that the brain interprets as color. Thus, colors exist only in our minds. This may sound strange, since we are used to seeing colors all the time, but consider the problem of describing a color to another person. All we can say is something like “this object is red,” but a color blind person has no idea of redness and is left uncomprehending. The reason we cannot describe colors is that they do not exist in nature. We cannot compare a color to any known attribute of the objects around us. The retina is the back part of the eye. It contains the light-sensitive (photoreceptor) cells which enable us to sense light and color. There are two types of photoreceptors, rods and cones. The rods are most sensitive to variations of light and dark, shape and movement. They contain one type of light-sensitive pigment and respond to even a few photons. They are therefore reponsible for black and white vision and for our dark-adapted (scotopic) vision. In dim light we use mainly our rods, which is why we don’t see colors under such conditions. There are about 120 million rods in the retina. The cones are much less numerous (there are only about 6–7 million of them and they are concentrated at the center of the retina, the fovea) and are much less sensitive to light (it takes hundreds of photons for a cone to respond). On the other hand, the cones are sensitive to wavelength, and it is they that send color information to the brain and are responsible for our color vision (more accurately, the cones are responsible for photopic vision, vision of the eye under well-lit conditions). There are three types of cones and they send three different types of stimuli to the brain. This is why any color specification requires three numbers (we perceive a three-dimensional color space). The three types of cones are sensitive to red, green, and blue, which is why these colors are a natural choice for primary colors. [More accurately, the three types of cones feature maximum sensitivity at wavelengths of about 420 nm (blue), 534 nm (Bluish-Green), and 564 nm (Yellowish-Green).] 7.10 JPEG 527 In addition to scotopic and photopic visions, there is also mesopic vision. This is a combination of photopic and scotopic visions in low light but not full darkness. Nothing in our world is perfect, and this includes people and other living beings. Color blindness in humans is not rare. It strikes about 8% of all males and about 0.5% of all females. It is caused by having just one or two types of cones (the other types may be completely missing or just weak). A color blind person with two types of cones can distinguish a limited range of colors, a range that requires only two numbers to specify a color. A person with only one type of cones senses even fewer colors and perceives a one-dimensional color space. With this in mind, can there be persons, animals, or aliens with four or more types of cones in their retina? The answer is yes, and such beings would perceive color spaces of more than three dimensions and would sense more colors that we do! This surprising conclusion stems from the fact that colors are not attributes of objects and exist only in our minds. The famous “blind spot” also deserves mentioning. This is a point where the retina does not have any photoreceptors. Any image in this region is not sensed by the eye and is invisible. The blind spot is the point where the optic nerves come together and exit the eye on their way to the brain. To find your blind spot, look at this image, close your left eye, and hold the page about 20 inches (or 50 cm) from you. Look at the square with your right eye. Slowly bring the page closer while looking at the square. At a certain distance, the cross will disappear. This occurs when the cross falls on the blind spot of your retina. Reverse the process. Close your right eye and look at the cross with your left eye. Bring the image slowly closer to you and the square will disappear. Each of the photoreceptors sends a light sensation to the brain that’s essentially a pixel, and the brain combines these pixels to a continuous image. Thus, the human eye is similar to a digital camera. Once this is understood, we naturally want to compare the resolution of the eye to that of a modern digital camera. Current (2009) digital cameras feature from 1–2 Mpixels (for a cell phone camera) to about ten Mpixels (for a goodquality camera). G R B 0 .02 .04 .06 .08 .10 .12 .14 .16 .18 .20 400 440 480 520 560 600 640 680 wavelength (nm) Fraction of light absorbed by each type of cone Thus, the eye features a much higher resolu- Figure 7.62: Sensitivity of the Cones. tion, but its effective resolution is even higher if we consider that the eye can move and refocus itself about three to four times a second. This means that in a single second, the eye can sense and send to the brain about half a billion pixels. Assuming that our camera takes a snapshot once a second, the ratio of the resolutions is about 100. 528 7. Image Compression Certain colors—such as red, orange, and yellow—are psychologically associated with heat. They are considered warm and cause a picture to appear larger and closer than it actually is. Other colors—such as blue, violet, and green—are associated with cool things (air, sky, water, ice) and are therefore called cool colors. They cause a picture to look smaller and farther away. 7.10.2 DCT The general concept of a transform is discussed in Section 7.6. The discrete cosine transform is discussed in much detail in Section 7.8. Other examples of important transforms are the Fourier transform (Section 8.1) and the wavelet transform (Chapter 8). Both have applications in many areas and also have discrete versions (DFT and DWT). The JPEG committee elected to use the DCT because of its good performance, because it does not assume anything about the structure of the data (the DFT, for example, assumes that the data to be transformed is periodic), and because there are ways to speed it up (Section 7.8.5). The JPEG standard calls for applying the DCT not to the entire image but to data units (blocks) of 8Χ8 pixels. The reasons for this are (1) Applying DCT to large blocks involves many arithmetic operations and is therefore slow. Applying DCT to small data units is faster. (2) Experience shows that, in a continuous-tone image, correlations between pixels are short range. A pixel in such an image has a value (color component or shade of gray) that’s close to those of its near neighbors, but has nothing to do with the values of far neighbors. The JPEG DCT is therefore executed by Equation (7.16), duplicated here for n = 8 Gij = 1 4CiCj 7 x=0 7 y=0 pxy cos(2x + 1)i. 16 cos(2y + 1)j. 16 , where Cf = 1 .2 , f = 0, 1, f>0, and 0 . i, j . 7. (7.16) The DCT is JPEG’s key to lossy compression. The unimportant image information is reduced or removed by quantizing the 64 DCT coefficients, especially the ones located toward the lower-right. If the pixels of the image are correlated, quantization does not degrade the image quality much. For best results, each of the 64 coefficients is quantized by dividing it by a different quantization coefficient (QC). All 64 QCs are parameters that can be controlled, in principle, by the user (Section 7.10.3). The JPEG decoder works by computing the inverse DCT (IDCT), Equation (7.17), duplicated here for n = 8 pxy = 1 4 7 i=0 7 j=0 CiCjGij cos(2x + 1)i. 16 cos(2y + 1)j. 16 , where Cf = 1 .2 , f = 0; 1, f>0. (7.17) 7.10 JPEG 529 It takes the 64 quantized DCT coefficients and calculates 64 pixels pxy. If the QCs are the right ones, the new 64 pixels will be very similar to the original ones. Mathematically, the DCT is a one-to-one mapping of 64-point vectors from the image domain to the frequency domain. The IDCT is the reverse mapping. If the DCT and IDCT could be calculated with infinite precision and if the DCT coefficients were not quantized, the original 64 pixels would be exactly reconstructed. 7.10.3 Quantization After each 8Χ8 data unit of DCT coefficients Gij is computed, it is quantized. This is the step where information is lost (except for some unavoidable loss because of finite precision calculations in other steps). Each number in the DCT coefficients matrix is divided by the corresponding number from the particular “quantization table” used, and the result is rounded to the nearest integer. As has already been mentioned, three such tables are needed, for the three color components. The JPEG standard allows for up to four tables, and the user can select any of the four for quantizing each color component. The 64 numbers that constitute each quantization table are all JPEG parameters. In principle, they can all be specified and fine-tuned by the user for maximum compression. In practice, few users have the patience or expertise to experiment with so many parameters, so JPEG software normally uses the following two approaches: 1. Default quantization tables. Two such tables, for the luminance (grayscale) and the chrominance components, are the result of many experiments performed by the JPEG committee. They are included in the JPEG standard and are reproduced here as Table 7.63. It is easy to see how the QCs in the table generally grow as we move from the upper left corner to the bottom right corner. This is how JPEG reduces the DCT coefficients with high spatial frequencies. 2. A simple quantization table Q is computed, based on one parameter R specified by the user. A simple expression such as Qij = 1+(i + j) Χ R guarantees that QCs start small at the upper-left corner and get bigger toward the lower-right corner. Table 7.64 shows an example of such a table with R = 2. 16 11 10 16 24 40 51 61 12 12 14 19 26 58 60 55 14 13 16 24 40 57 69 56 14 17 22 29 51 87 80 62 18 22 37 56 68 109 103 77 24 35 55 64 81 104 113 92 49 64 78 87 103 121 120 101 72 92 95 98 112 100 103 99 17 18 24 47 99 99 99 99 18 21 26 66 99 99 99 99 24 26 56 99 99 99 99 99 47 66 99 99 99 99 99 99 99 99 99 99 99 99 99 99 99 99 99 99 99 99 99 99 99 99 99 99 99 99 99 99 99 99 99 99 99 99 99 99 Luminance Chrominance Table 7.63: Recommended Quantization Tables. If the quantization is done correctly, very few nonzero numbers will be left in the DCT coefficients matrix, and they will typically be concentrated in the upper-left region. These numbers are the output of JPEG, but they are further compressed before being 530 7. Image Compression 1 3 5 7 9 11 13 15 3 5 7 9 11 13 15 17 5 7 9 11 13 15 17 19 7 9 11 13 15 17 19 21 9 11 13 15 17 19 21 23 11 13 15 17 19 21 23 25 13 15 17 19 21 23 25 27 15 17 19 21 23 25 27 29 Table 7.64: The Quantization Table 1 + (i + j) Χ 2. written on the output stream. In the JPEG literature this compression is called “entropy coding,” and Section 7.10.4 shows in detail how it is done. Three techniques are used by entropy coding to compress the 8 Χ 8 matrix of integers: 1. The 64 numbers are collected by scanning the matrix in zigzags (Figure 1.8b). This produces a string of 64 numbers that starts with some nonzeros and typically ends with many consecutive zeros. Only the nonzero numbers are output (after further compressing them) and are followed by a special end-of block (EOB) code. This way there is no need to output the trailing zeros (we can say that the EOB is the run-length encoding of all the trailing zeros). The interested reader should also consult Section 11.5 for other methods to compress binary strings with many consecutive zeros.  Exercise 7.16: Propose a practical way to write a loop that traverses an 8 Χ 8 matrix in zigzag. 2. The nonzero numbers are compressed using Huffman coding (Section 7.10.4). 3. The first of those numbers (the DC coefficient, page 471) is treated differently from the others (the AC coefficients). She had just succeeded in curving it down into a graceful zigzag, and was going to dive in among the leaves, which she found to be nothing but the tops of the trees under which she had been wandering, when a sharp hiss made her draw back in a hurry. —Lewis Carroll, Alice in Wonderland (1865) 7.10.4 Coding We first discuss point 3 above. Each 8Χ8 matrix of quantized DCT coefficients contains one DC coefficient [at position (0, 0), the top left corner] and 63 AC coefficients. The DC coefficient is a measure of the average value of the 64 original pixels, constituting the data unit. Experience shows that in a continuous-tone image, adjacent data units of pixels are normally correlated in the sense that the average values of the pixels in adjacent data units are close. We already know that the DC coefficient of a data unit is a multiple of the average of the 64 pixels constituting the unit. This implies that the DC coefficients of adjacent data units don’t differ much. JPEG outputs the first one (encoded), followed by differences (also encoded) of the DC coefficients of consecutive data units. The concept of differencing is discussed in Section 1.3.1. 7.10 JPEG 531 Example: If the first three 8Χ8 data units of an image have quantized DC coefficients of 1118, 1114, and 1119, then the JPEG output for the first data unit is 1118 (Huffman encoded, see below) followed by the 63 (encoded) AC coefficients of that data unit. The output for the second data unit will be 1114 - 1118 = -4 (also Huffman encoded), followed by the 63 (encoded) AC coefficients of that data unit, and the output for the third data unit will be 1119 - 1114 = 5 (also Huffman encoded), again followed by the 63 (encoded) AC coefficients of that data unit. This way of handling the DC coefficients is worth the extra trouble, because the differences are small. Coding the DC differences is done with Table 7.65, so first here are a few words about this table. Each row has a row number (on the left), the unary code for the row (on the right), and several columns in between. Each row contains greater numbers (and also more numbers) than its predecessor but not the numbers contained in previous rows. Row i contains the range of integers [-(2i-1), +(2i-1)] but is missing the middle range [-(2i-1 - 1), +(2i-1 - 1)]. Thus, the rows get very long, which means that a simple two-dimensional array is not a good data structure for this table. In fact, there is no need to store these integers in a data structure, since the program can figure out where in the table any given integer x is supposed to reside by analyzing the bits of x. The first DC coefficient to be encoded in our example is 1118. It resides in row 11 column 930 of the table (column numbering starts at zero), so it is encoded as 111111111110|01110100010 (the unary code for row 11, followed by the 11-bit binary value of 930). The second DC difference is -4. It resides in row 3 column 3 of Table 7.65, so it is encoded as 1110|011 (the unary code for row 3, followed by the 3-bit binary value of 3).  Exercise 7.17: How is the third DC difference, 5, encoded? Point 2 above has to do with the precise way the 63 AC coefficients of a data unit are compressed. It uses a combination of RLE and either Huffman or arithmetic coding. The idea is that the sequence of AC coefficients normally contains just a few nonzero numbers, with runs of zeros between them, and with a long run of trailing zeros. For each nonzero number x, the encoder (1) finds the number Z of consecutive zeros preceding x; (2) finds x in Table 7.65 and prepares its row and column numbers (R and C); (3) the pair (R, Z) [that’s (R, Z), not (R, C)] is used as row and column numbers for Table 7.68; and (4) the Huffman code found in that position in the table is concatenated to C (where C is written as an R-bit number) and the result is (finally) the code emitted by the JPEG encoder for the AC coefficient x and all the consecutive zeros preceding it. The Huffman codes in Table 7.68 are not the ones recommended by the JPEG standard. The standard recommends the use of Tables 7.66 and 7.67 and says that up to four Huffman code tables can be used by a JPEG codec, except that the baseline mode can use only two such tables. The actual codes in Table 7.68 are thus arbitrary. The reader should notice the EOB code at position (0, 0) and the ZRL code at position (0, 15). The former indicates end-of-block, and the latter is the code emitted for 15 consecutive zeros when the number of consecutive zeros exceeds 15. These codes are the ones recommended for the luminance AC coefficients of Table 7.66. The EOB and ZRL codes recommended for the chrominance AC coefficients of Table 7.67 are 00 and 1111111010, respectively. 532 7. Image Compression 0: 0 0 1: -1 1 10 2: -3 -2 2 3 110 3: -7 -6 -5 -4 4 5 6 7 1110 4: -15 -14 . . . -9 -8 8 9 10 . . . 15 11110 5: -31 -30 -29 . . . -17 -16 16 17 . . . 31 111110 6: -63 -62 -61 . . . -33 -32 32 33 . . . 63 1111110 7: -127 -126 -125 . . . -65 -64 64 65 . . . 127 11111110 ... ... 14: -16383 -16382 -16381 . . . -8193 -8192 8192 8193 . . . 16383 111111111111110 15: -32767 -32766 -32765 . . . -16385 -16384 16384 16385 . . . 32767 1111111111111110 16: 32768 1111111111111111 Table 7.65: Coding the Differences of DC Coefficients. As an example consider the sequence 1118, 2, 0,-2, 0, . . . , 0    13 ,-1, 0, . . . . The first AC coefficient 2 has no zeros preceding it, so Z = 0. It is found in Table 7.65 in row 2, column 2, so R = 2 and C = 2. The Huffman code in position (R, Z) = (2, 0) of Table 7.68 is 01, so the final code emitted for 2 is 01|10. The next nonzero coefficient, -2, has one zero preceding it, so Z = 1. It is found in Table 7.65 in row 2, column 1, so R = 2 and C = 1. The Huffman code in position (R, Z) = (2, 1) of Table 7.68 is 11011, so the final code emitted for 2 is 11011|01.  Exercise 7.18: What code is emitted for the last nonzero AC coefficient, -1? Finally, the sequence of trailing zeros is encoded as 1010 (EOB), so the output for the above sequence of AC coefficients is 01101101110111010101010. We saw earlier that the DC coefficient is encoded as 111111111110|1110100010, so the final output for the entire 64-pixel data unit is the 46-bit number 1111111111100111010001001101101110111010101010. These 46 bits encode one color component of the 64 pixels of a data unit. Let’s assume that the other two color components are also encoded into 46-bit numbers. If each pixel originally consists of 24 bits, then this corresponds to a compression factor of 64 Χ 24/(46 Χ 3) . 11.13; very impressive! (Notice that the DC coefficient of 1118 has contributed 23 of the 46 bits. Subsequent data units code differences of their DC coefficient, which may take fewer than 10 bits instead of 23. They may feature much higher compression factors as a result.) The same tables (Tables 7.65 and 7.68) used by the encoder should, of course, be used by the decoder. The tables may be predefined and used by a JPEG codec as defaults, or they may be specifically calculated for a given image in a special pass preceding the actual compression. The JPEG standard does not specify any code tables, so any JPEG codec must use its own. Readers who feel that this coding scheme is complex should take a look at the much more complex CAVLC coding method that is employed by H.264 (Section 9.9) to encode a similar sequence of 8Χ8 DCT transform coefficients. 7.10 JPEG 533 R Z 1 2 3 4 5 6 7 8 9 A 0 00 01 100 1011 11010 1111000 11111000 1111110110 1111111110000010 1111111110000011 1 1100 11011 11110001 111110110 11111110110 1111111110000100 1111111110000101 1111111110000110 1111111110000111 1111111110001000 2 11100 11111001 1111110111 111111110100 111111110001001 111111110001010 111111110001011 111111110001100 111111110001101 111111110001110 3 111010 111110111 111111110101 1111111110001111 1111111110010000 1111111110010001 1111111110010010 1111111110010011 1111111110010100 1111111110010101 4 111011 1111111000 1111111110010110 1111111110010111 1111111110011000 1111111110011001 1111111110011010 1111111110011011 1111111110011100 1111111110011101 5 1111010 11111110111 1111111110011110 1111111110011111 1111111110100000 1111111110100001 1111111110100010 1111111110100011 1111111110100100 1111111110100101 6 1111011 111111110110 1111111110100110 1111111110100111 1111111110101000 1111111110101001 1111111110101010 1111111110101011 1111111110101100 1111111110101101 7 11111010 111111110111 1111111110101110 1111111110101111 1111111110110000 1111111110110001 1111111110110010 1111111110110011 1111111110110100 1111111110110101 8 111111000 111111111000000 1111111110110110 1111111110110111 1111111110111000 1111111110111001 1111111110111010 1111111110111011 1111111110111100 1111111110111101 9 111111001 1111111110111110 1111111110111111 1111111111000000 1111111111000001 1111111111000010 1111111111000011 1111111111000100 1111111111000101 1111111111000110 A 111111010 1111111111000111 1111111111001000 1111111111001001 1111111111001010 1111111111001011 1111111111001100 1111111111001101 1111111111001110 1111111111001111 B 1111111001 1111111111010000 1111111111010001 1111111111010010 1111111111010011 1111111111010100 1111111111010101 1111111111010110 1111111111010111 1111111111011000 C 1111111010 1111111111011001 1111111111011010 1111111111011011 1111111111011100 1111111111011101 1111111111011110 1111111111011111 1111111111100000 1111111111100001 D 11111111000 1111111111100010 1111111111100011 1111111111100100 1111111111100101 1111111111100110 1111111111100111 1111111111101000 1111111111101001 1111111111101010 E 1111111111101011 1111111111101100 1111111111101101 1111111111101110 1111111111101111 1111111111110000 1111111111110001 1111111111110010 1111111111110011 1111111111110100 F 11111111001 1111111111110101 1111111111110110 1111111111110111 1111111111111000 1111111111111001 1111111111111010 1111111111111011 1111111111111101 1111111111111110 Table 7.66: Recommended Huffman Codes for Luminance AC Coefficients. 534 7. Image Compression R Z 1 2 3 4 5 6 7 8 9 A 0 01 100 1010 11000 11001 111000 1111000 111110100 1111110110 111111110100 1 1011 111001 11110110 111110101 11111110110 111111110101 111111110001000 111111110001001 111111110001010 111111110001011 2 11010 11110111 1111110111 111111110110 111111111000010 1111111110001100 1111111110001101 1111111110001110 1111111110001111 1111111110010000 3 11011 11111000 1111111000 111111110111 1111111110010001 1111111110010010 1111111110010011 1111111110010100 1111111110010101 1111111110010110 4 111010 111110110 1111111110010111 1111111110011000 1111111110011001 1111111110011010 1111111110011011 1111111110011100 1111111110011101 1111111110011110 5 111011 1111111001 1111111110011111 1111111110100000 1111111110100001 1111111110100010 1111111110100011 1111111110100100 1111111110100101 1111111110100110 6 1111001 11111110111 1111111110100111 1111111110101000 1111111110101001 1111111110101010 1111111110101011 1111111110101100 1111111110101101 1111111110101110 7 1111010 11111111000 1111111110101111 1111111110110000 1111111110110001 1111111110110010 1111111110110011 1111111110110100 1111111110110101 1111111110110110 8 11111001 1111111110110111 1111111110111000 1111111110111001 1111111110111010 1111111110111011 1111111110111100 1111111110111101 1111111110111110 1111111110111111 9 111110111 1111111111000000 1111111111000001 1111111111000010 1111111111000011 1111111111000100 1111111111000101 1111111111000110 1111111111000111 1111111111001000 A 111111000 1111111111001001 1111111111001010 1111111111001011 1111111111001100 1111111111001101 1111111111001110 1111111111001111 1111111111010000 1111111111010001 B 111111001 1111111111010010 1111111111010011 1111111111010100 1111111111010101 1111111111010110 1111111111010111 1111111111011000 1111111111011001 1111111111011010 C 111111010 1111111111011011 1111111111011100 1111111111011101 1111111111011110 1111111111011111 1111111111100000 1111111111100001 1111111111100010 1111111111100011 D 11111111001 1111111111100100 1111111111100101 1111111111100110 1111111111100111 1111111111101000 1111111111101001 1111111111101010 1111111111101011 1111111111101100 E 11111111100000 1111111111101101 1111111111101110 1111111111101111 1111111111110000 1111111111110001 1111111111110010 1111111111110011 1111111111110100 1111111111110101 F 111111111000011 111111111010110 1111111111110111 1111111111111000 1111111111111001 1111111111111010 1111111111111011 1111111111111100 1111111111111101 1111111111111110 Table 7.67: Recommended Huffman Codes for Chrominance AC Coefficients. 7.10 JPEG 535 R Z: 0 1 . . . 15 0: 1010 11111111001(ZRL) 1: 00 1100 . . . 1111111111110101 2: 01 11011 . . . 1111111111110110 3: 100 1111001 . . . 1111111111110111 4: 1011 111110110 . . . 1111111111111000 5: 11010 11111110110 . . . 1111111111111001 ... ... Table 7.68: Coding AC Coefficients. Some JPEG variants use a particular version of arithmetic coding, called the QM coder (Section 5.11), that is specified in the JPEG standard. This version of arithmetic coding is adaptive, so it does not need Tables 7.65 and 7.68. It adapts its behavior to the image statistics as it goes along. Using arithmetic coding may produce 5–10% better compression than Huffman for a typical continuous-tone image. However, it is more complex to implement than Huffman coding, so in practice it is rare to find a JPEG codec that uses it. 7.10.5 Lossless Mode The lossless mode of JPEG uses differencing (Section 1.3.1) to reduce the values of pixels before they are compressed. This particular form of differencing is called predicting. The values of some near neighbors of a pixel are subtracted from the pixel to get a small number, which is then compressed further using Huffman or arithmetic coding. Figure 7.69a shows a pixel X and three neighbor pixels A, B, and C. Figure 7.69b shows eight possible ways (predictions) to combine the values of the three neighbors. In the lossless mode, the user can select one of these predictions, and the encoder then uses it to combine the three neighbor pixels and subtract the combination from the value of X. The result is normally a small number, which is then entropy-coded in a way very similar to that described for the DC coefficient in Section 7.10.4. Predictor 0 is used only in the hierarchical mode of JPEG. Predictors 1, 2, and 3 are called one-dimensional. Predictors 4, 5, 6, and 7 are two-dimensional. It should be noted that the lossless mode of JPEG has never been very successful. It produces typical compression factors of 2, and is therefore inferior to other lossless image compression methods. Because of this, many JPEG implementations do not even implement this mode. Even the lossy (baseline) mode of JPEG does not perform well when asked to limit the amount of loss to a minimum. As a result, some JPEG implementations do not allow parameter settings that result in minimum loss. The strength of JPEG is in its ability to generate highly compressed images that when decompressed are indistinguishable from the original. Recognizing this, the ISO has decided to come up with another standard for lossless compression of continuous-tone images. This standard is now commonly known as JPEG-LS and is described in Section 7.11. 536 7. Image Compression C B A X Selection value Prediction 0 no prediction 1 A 2 B 3 C 4 A+B- C 5 A + ((B - C)/2) 6 B + ((A - C)/2) 7 (A+B)/2 (a) (b) Figure 7.69: Pixel Prediction in the Lossless Mode. 7.10.6 The Compressed File A JPEG encoder outputs a compressed file that includes parameters, markers, and the compressed data units. The parameters are either four bits (these always come in pairs), one byte, or two bytes long. The markers serve to identify the various parts of the file. Each is two bytes long, where the first byte is X’FF’ and the second one is not 0 or X’FF’. A marker may be preceded by a number of bytes with X’FF’. Table 7.71 lists all the JPEG markers (the first four groups are start-of-frame markers). The compressed data units are combined into MCUs (minimal coded unit), where an MCU is either a single data unit (in the noninterleaved mode) or three data units from the three image components (in the interleaved mode). Compressed [Tables] [Tables] [DNL segment] SOI Frame Frame EOI MCU MCU MCU MCU MCU MCU image Frame header Segment0 Segmentlast Frame header Scan1 Scan ECS0 [RST0] ECSlast-1 [RSTlast-1] ECSlast [Scan2] [Scanlast] Figure 7.70: JPEG File Format. 7.10 JPEG 537 Value Name Description Nondifferential, Huffman coding FFC0 SOF0 Baseline DCT FFC1 SOF1 Extended sequential DCT FFC2 SOF2 Progressive DCT FFC3 SOF3 Lossless (sequential) Differential, Huffman coding FFC5 SOF5 Differential sequential DCT FFC6 SOF6 Differential progressive DCT FFC7 SOF7 Differential lossless (sequential) Nondifferential, arithmetic coding FFC8 JPG Reserved for extensions FFC9 SOF9 Extended sequential DCT FFCA SOF10 Progressive DCT FFCB SOF11 Lossless (sequential) Differential, arithmetic coding FFCD SOF13 Differential sequential DCT FFCE SOF14 Differential progressive DCT FFCF SOF15 Differential lossless (sequential) Huffman table specification FFC4 DHT Define Huffman table Arithmetic coding conditioning specification FFCC DAC Define arith coding conditioning(s) Restart interval termination FFD0–FFD7 RSTm Restart with modulo 8 count m Other markers FFD8 SOI Start of image FFD9 EOI End of image FFDA SOS Start of scan FFDB DQT Define quantization table(s) FFDC DNL Define number of lines FFDD DRI Define restart interval FFDE DHP Define hierarchical progression FFDF EXP Expand reference component(s) FFE0–FFEF APPn Reserved for application segments FFF0–FFFD JPGn Reserved for JPEG extensions FFFE COM Comment Reserved markers FF01 TEM For temporary private use FF02–FFBF RES Reserved Table 7.71: JPEG Markers. 538 7. Image Compression Figure 7.70 shows the main parts of the JPEG compressed file (parts in square brackets are optional). The file starts with the SOI marker and ends with the EOI marker. In between these markers, the compressed image is organized in frames. In the hierarchical mode there are several frames, and in all other modes there is only one frame. In each frame the image information is contained in one or more scans, but the frame also contains a header and optional tables (which, in turn, may include markers). The first scan may be followed by an optional DNL segment (define number of lines), which starts with the DNL marker and contains the number of lines in the image that’s represented by the frame. A scan starts with optional tables, followed by the scan header, followed by several entropy-coded segments (ECS), which are separated by (optional) restart markers (RST). Each ECS contains one or more MCUs, where an MCU is, as explained earlier, either a single data unit or three such units. See Section 7.40 for an image compression method that combines JPEG with a fractal-based method. 7.10.7 JFIF It has been mentioned earlier that JPEG is a compression method, not a graphics file format, which is why it does not specify image features such as pixel aspect ratio, color space, or interleaving of bitmap rows. This is where JFIF comes in. JFIF (JPEG File Interchange Format) is a graphics file format that makes it possible to exchange JPEG-compressed images between computers. The main features of JFIF are the use of the YCbCr triple-component color space for color images (only one component for grayscale images) and the use of a marker to specify features missing from JPEG, such as image resolution, aspect ratio, and features that are application-specific. The JFIF marker (called the APP0 marker) starts with the zero-terminated string JFIF. Following this, there is pixel information and other specifications (see below). Following this, there may be additional segments specifying JFIF extensions. A JFIF extension contains more platform-specific information about the image. Each extension starts with the zero-terminated string JFXX, followed by a 1-byte code identifying the extension. An extension may contain application-specific information, in which case it starts with a different string, not JFIF or JFXX but something that identifies the specific application or its maker. The format of the first segment of an APP0 marker is as follows: 1. APP0 marker (4 bytes): FFD8FFE0. 2. Length (2 bytes): Total length of marker, including the 2 bytes of the “length” field but excluding the APP0 marker itself (field 1). 3. Identifier (5 bytes): 4A4649460016. This is the JFIF string that identifies the APP0 marker. 4. Version (2 bytes): Example: 010216 specifies version 1.02. 5. Units (1 byte): Units for the X and Y densities. 0 means no units; the Xdensity and Ydensity fields specify the pixel aspect ratio. 1 means that Xdensity and Ydensity are dots per inch, 2, that they are dots per cm. 6. Xdensity (2 bytes), Ydensity (2 bytes): Horizontal and vertical pixel densities (both should be nonzero). 7. Xthumbnail (1 byte), Ythumbnail (1 byte): Thumbnail horizontal and vertical pixel counts. 7.10 JPEG 539 8. (RGB)n (3n bytes): Packed (24-bit) RGB values for the thumbnail pixels. n = XthumbnailΧYthumbnail. The syntax of the JFIF extension APP0 marker segment is as follows: 1. APP0 marker. 2. Length (2 bytes): Total length of marker, including the 2 bytes of the “length” field but excluding the APP0 marker itself (field 1). 3. Identifier (5 bytes): 4A4658580016 This is the JFXX string identifying an extension. 4. Extension code (1 byte): 1016 = Thumbnail coded using JPEG. 1116 = Thumbnail coded using 1 byte/pixel (monochromatic). 1316 = Thumbnail coded using 3 bytes/pixel (eight colors). 5. Extension data (variable): This field depends on the particular extension. JFIF is the technical name for the image format better (but inaccurately) known as JPEG. This term is used only when the difference between the Image Format and the Image Compression is crucial. Strictly speaking, however, JPEG does not define an Image Format, and therefore in most cases it would be more precise to speak of JFIF rather than JPEG. Another Image Format for JPEG is SPIFF defined by the JPEG standard itself, but JFIF is much more widespread than SPIFF. —Erik Wilde, WWW Online Glossary 7.10.8 HD photo and JPEG XR HD Photo is a complete specification (i.e., an algorithm and a file format) for the compression of continuous-tone photographic images. This method (formerly known as Windows Media Photo and also as Photon) was developed and patented by Microsoft as part of its Windows Media family. It supports both lossy and lossless compression, and is the standard image format for Microsoft’s XPS documents. Microsoft official implementations of HD photo are part of the Windows Vista operating system and are also available for Windows XP. The Joint Photographic Experts Group (the developer of JPEG) has decided to adopt HD photo as a JPEG standard, to be titled JPEG XR (extended range). The formal description of JPEG XR is ISO/IEC 29199-2. An important feature of HD Photo is the use of integer operations only (even division is not required) in both encoder and decoder. The method supports bi-level, RGB, CMYK, and even n-channel color representation, with up to 16-bit unsigned integer representation, or up to 32-bit fixed-point or floating-point representation. The original color scheme is transformed to an internal color space and quantization is applied to achieve lossy compression. The algorithm has features for decoding just part of an image. Certain operations such as cropping, downsampling, horizontal or vertical flips, and 90. rotations can also be performed without decoding the entire image. An HD Photo file has a format similar to TIFF (with all the advantages and limitations of this format, such as a maximum 4 GB file size). The file contains Image File Directory (IFD) tags, the image data (self-contained), optional alpha channel data (compressed as a separate image record), HD Photo metadata, optional XMP metadata, and optional Exif metadata, in IFD tags. The compression algorithm of HD Photo is based on JPEG. The input image is converted to a luminance chrominance color space, the chroma components are optionally 540 7. Image Compression subsampled to increase compression, each of the three color components is partitioned into fixed-size blocks which are transformed into the frequency domain, and the frequency coefficients are quantized and entropy coded. The main differences between the original JPEG and HD photo are the following: JPEG supports 8-bit pixels (with some implementations also handling 12-bit pixels), while HD Photo supports pixels of up to 32 bits. HD Photo also includes features for the lossless compression of floating-point pixels. The color transformation of JPEG (from RGB to YCbCr) is reversible in principle, but in practice involves small losses of data because of roundoff errors. In contrast, HD Photo employs a lossless color transformation, that for RGB is given by Co = B - R, Cg = R - G + $Co 2 %, Y = G + Cg 2 !. JPEG uses blocks of 8 Χ 8 pixels for its DCT, while HD Photo also uses blocks of 4Χ4, 2Χ4, and 2Χ2 pixels. These block sizes can be selected by the encoder for image regions with lower-than-normal pixel correlation. The DCT is also reversible in principle and slightly lossy in practice because of the limited precision of computer arithmetic. The transform used by HD Photo is called the photo core transform (PCT). It is similar to the Walsh-Hadamard transform (Section 7.7.2), but is lossless because it employs only integers. Before applying the PCT, HD Photo offers an optional, lossless prefiltering step that has a visual blurring effect. This step filters 4 Χ 4 blocks that are offset by two pixels in each direction from the PCT blocks. Experiments indicate that this type of filtering reduces the annoying block-boundary JPEG artifacts that occur mostly at low bitrates. This step is normally skipped at high bitrates, where such artifacts are negligible or nonexistent. Section 7.10.4 explains how JPEG encodes the DCT coefficients (specifically, how the DC coefficients are encoded by left-prediction). In HD Photo, blocks are grouped into macroblocks of 16 Χ 16 pixels, and the DC components from each macroblock are themselves frequency transformed. Thus, HD photo has to encode three types of coefficients: the macroblock DC coefficients (called DC), macroblock-level AC coefficients (termed lowpass), and lower-level AC coefficients (referred to as AC). The encoding of these three types is far more complex in HD Photo than JPEG encoding. It consists of (1) a complex DC prediction algorithm, (2) an adaptive coefficient reordering (instead of simple zigzag ordering), followed by (3) a form of adaptive Huffman coding for the coefficients themselves. There is a single quantization coefficient for the DC transform coefficients in each color plane. The lowpass and the AC quantization coefficients, however, can vary from macroblock to macroblock. 7.11 JPEG-LS 541 HD Photo produces lossless compression simply by setting all its quantization coefficients to 1. Because of the use of integer transform and special color space transformation, there are no other losses in HD photo. This is in contrast with JPEG, whose lossless mode is inefficient and often is not even implemented. 7.11 JPEG-LS As has been mentioned in Section 7.10.5, the lossless mode of JPEG is inefficient and often is not even implemented. As a result, the ISO, in cooperation with the IEC, has decided to develop a new standard for the lossless (or near-lossless) compression of continuous-tone images. The result is recommendation ISO/IEC CD 14495, popularly known as JPEG-LS. The principles of this method are described here, but it should be noted that it is not simply an extension or a modification of JPEG. This is a new method, designed to be simple and fast. It does not use the DCT, does not employ arithmetic coding, and uses quantization in a restricted way, and only in its near-lossless option. JPEG-LS is based on ideas developed in [Weinberger et al. 96 and 00] and for their LOCO-I compression method. JPEG-LS examines several of the previously-seen neighbors of the current pixel, uses them as the context of the pixel, uses the context to predict the pixel and to select a probability distribution out of several such distributions, and uses that distribution to encode the prediction error with a special Golomb code. There is also a run mode, where the length of a run of identical pixels is encoded. The context used to predict the current pixel x is shown in Figure 7.72. The encoder examines the context pixels and decides whether to encode the current pixel x in the run mode or in the regular mode. If the context suggests that the pixels y, z,. . . following the current pixel are likely to be identical, the encoder selects the run mode. Otherwise, it selects the regular mode. In the near-lossless mode the decision is slightly different. If the context suggests that the pixels following the current pixel are likely to be almost identical (within the tolerance parameter NEAR), the encoder selects the run mode. Otherwise, it selects the regular mode. The rest of the encoding process depends on the mode selected. c b d a x y z Figure 7.72: Context for Predicting x. In the regular mode, the encoder uses the values of context pixels a, b, and c to predict pixel x, and subtracts the prediction from x to obtain the prediction error, denoted by Errval. This error is then corrected by a term that depends on the context (this correction is done to compensate for systematic biases in the prediction), and encoded with a Golomb code. The Golomb coding depends on all four pixels of the context and also on prediction errors that were previously encoded for the same context 542 7. Image Compression (this information is stored in arrays A and N, mentioned in Section 7.11.1). If nearlossless compression is used, the error is quantized before it is encoded. In the run mode, the encoder starts at the current pixel x and finds the longest run of pixels that are identical to context pixel a. The encoder does not extend this run beyond the end of the current image row. Since all the pixels in the run are identical to a (and a is already known to the decoder) only the length of the run needs be encoded, and this is done with a 32-entry array denoted by J (Section 7.11.1). If near-lossless compression is used, the encoder selects a run of pixels that are close to a within the tolerance parameter NEAR. The decoder is not substantially different from the encoder, so JPEG-LS is a nearly symmetric compression method. The compressed stream contains data segments (with the Golomb codes and the encoded run lengths), marker segments (with information needed by the decoder), and markers (some of the reserved markers of JPEG are used). A marker is a byte of all ones followed by a special code, signaling the start of a new segment. If a marker is followed by a byte whose most significant bit is 0, that byte is the start of a marker segment. Otherwise, that byte starts a data segment. 7.11.1 The Encoder JPEG-LS is normally used as a lossless compression method. In this case, the reconstructed value of a pixel is identical to its original value. In the near lossless mode, the original and the reconstructed values may differ. In every case we denote the reconstructed value of a pixel p by Rp. When the top row of an image is encoded, context pixels b, c, and d are not available and should therefore be considered zero. If the current pixel is located at the start or the end of an image row, either a and c, or d are not available. In such a case, the encoder uses for a or d the reconstructed value Rb of b (or zero if this is the top row), and for c the reconstructed value that was assigned to a when the first pixel of the previous line was encoded. This means that the encoder has to do part of the decoder’s job and has to figure out the reconstructed values of certain pixels. The first step in determining the context is to calculate the three gradient values D1 = Rd - Rb, D2 = Rb - Rc, D3 = Rc - Ra. If the three values are zero (or, for near-lossless, if their absolute values are less than or equal to the tolerance parameter NEAR), the encoder selects the run mode, where it looks for the longest run of pixels identical to Ra. Step 2 compares the three gradients Di to certain parameters and calculates three region numbers Qi according to certain rules (not shown here). Each region number Qi can take one of the nine integer values in the interval [-4, +4], so there are 9Χ9Χ9 = 729 different region numbers. The third step maps the region numbers Qi to an integer Q in the interval [0, 364]. The tuple (0, 0, 0) is mapped to 0, and the 728 remaining tuples are mapped to [1, 728/2 = 364], such that (a, b, c) and (-a,-b,-c) are mapped to the same value. The details of this calculation are not specified by the JPEG-LS standard, and the encoder can do it in any way it chooses. The integer Q becomes the context for the current pixel x. It is used to index arrays A and N in Figure 7.76. After determining the context Q, the encoder predicts pixel x in two steps. The first step calculates the prediction Px based on edge rules, as shown in Figure 7.73. 7.11 JPEG-LS 543 The second step corrects the prediction, as shown in Figure 7.74, based on the quantity SIGN (determined from the signs of the three regions Qi), the correction values C[Q] (derived from the bias and not discussed here), and parameter MAXVAL. if(Rc>=max(Ra,Rb)) Px=min(Ra,Rb); else if(Rc<=min(Ra,Rb)) Px=max(Ra,Rb) else Px=Ra+Rb-Rc; endif; endif; Figure 7.73: Edge Detecting. if(SIGN=+1) Px=Px+C[Q] else Px=Px-C[Q] endif; if(Px>MAXVAL) Px=MAXVAL else if(Px<0) Px=0 endif; endif; Figure 7.74: Prediction Correcting. To understand the edge rules, let’s consider the case where b . a. In this case the edge rules select b as the prediction of x in many cases where a vertical edge exists in the image just left of the current pixel x. Similarly, a is selected as the prediction in many cases where a horizontal edge exists in the image just above x. If no edge is detected, the edge rules compute a prediction of a+b-c, and this has a simple geometric interpretation. If we interpret each pixel as a point in three-dimensional space, with the pixel’s intensity as its height, then the value a + b - c places the prediction Px on the same plane as pixels a, b, and c. Once the prediction Px is known, the encoder computes the prediction error Errval as the difference x - Px but reverses its sign if the quantity SIGN is negative. In the near-lossless mode the error is quantized, and the encoder uses it to compute the reconstructed value Rx of pixel x the way the decoder will do it in the future. The encoder needs this reconstructed value to encode future pixels. The basic quantization step is Errval ‹ Errval + NEAR 2ΧNEAR + 1 . It uses parameter NEAR, but it involves more details that are not shown here. The basic reconstruction step is Rx ‹ Px + SIGNΧErrvalΧ(2ΧNEAR + 1). The prediction error (after possibly being quantized) now goes through a range reduction (whose details are omitted here) and is finally ready for the important step of encoding. The Golomb code was introduced in Section 3.24, where its main parameter was denoted by b. JPEG-LS denotes this parameter by m. Once m has been selected, the Golomb code of the nonnegative integer n consists of two parts, the unary code of the integer part of n/m and the binary representation of n mod m. These codes are ideal for integers n that are distributed geometrically (i.e., when the probability of n is (1-r)rn, where 0 < r < 1). For any such geometric distribution there exists a value m such that the Golomb code based on it yields the shortest possible average code length. The special case where m is a power of 2 (m = 2k) leads to simple encoding/decoding operations. The code for n consists, in such a case, of the k least-significant bits of n, preceded by 544 7. Image Compression the unary code of the remaining most-significant bits of n. This particular Golomb code is denoted by G(k). As an example we compute the G(2) code of n = 19 = 100112. Since k = 2, m is 4. We start with the two least-significant bits, 11, of n. They equal the integer 3, which is also n mod m (3 = 19 mod 4). The remaining most-significant bits, 100, are also the integer 4, which is the integer part of the quotient n/m (19/4 = 4.75). The unary code of 4 is 00001, so the G(2) code of n = 19 is 00001|11. In practice, we always have a finite set of nonnegative integers, where the largest integer in the set is denoted by I. The maximum length of G(0) is I + 1, and since I can be large, it is desirable to limit the size of the Golomb code. This is done by the special Golomb code LG(k, glimit) which depends on the two parameters k and glimit. We first form a number q from the most significant bits of n. If q < glimit-log I-1, the LG(k, glimit) code is simply G(k). Otherwise, the unary code of glimit-log I-1 is prepared (i.e., glimit-log I-1 zeros followed by a single 1). This acts as an escape code and is followed by the binary representation of n - 1 in I bits. Our prediction errors are not necessarily positive. They are differences, so they can also be zero or negative, but the various Golomb codes were designed for nonnegative integers. This is why the prediction errors must be mapped to nonnegative values before they can be coded. This is done by MErrval = 2Errval, Errval . 0, 2|Errval| - 1, Errval<0. (7.25) This mapping interleaves negative and positive values in the sequence 0,-1, +1,-2, +2,-3, . . . . Table 7.75 lists some prediction errors, their mapped values, and their LG(2, 32) codes assuming an alphabet of size 256 (i.e., I = 255 and log I = 8). The next point to be discussed is how to determine the value of the Golomb code parameter k. This is done adaptively. Parameter k depends on the context, and the value of k for a context is updated each time a pixel with that context is found. The calculation of k can be expressed by the single C-language statement for (k=0; (N[Q]<>1; B[Q]=B[Q]>>1; N[Q]=N[Q]>>1 endif; N[Q]=N[Q]+1; Figure 7.76: Updating Arrays A, B, and N. 546 7. Image Compression value Ra of context pixel a. For near-lossless compression, pixels in the run must have values Ix that satisfy |Ix - Ra| . NEAR. A run is not allowed to continue beyond the end of the current image row. The length of the run is encoded (there is no need to encode the value of the run’s pixels, since it equals Ra), and if the run ends before the end of the current row, its encoded length is followed by the encoding of the pixel immediately following it (the pixel interrupting the run). The two main tasks of the encoder in this mode are (1) run scanning and runlength encoding and (2) run interruption coding. Run scanning is shown in Figure 7.77. Run-length encoding is shown in Figures 7.78 (for run segments of length rm) and 7.79 (for segments of length less than rm). Here are some of the details. RUNval=Ra; RUNcnt=0; while(abs(Ix-RUNval)<=NEAR) RUNcnt=RUNcnt+1; Rx=RUNval; if(EOLine=1) break else GetNextSample() endif; endwhile; Figure 7.77: Run Scanning. while(RUNcnt>=(1<0) RUNindex=RUNindex-1; endif; else if(RUNcnt>0) AppendToBitStream(1,1); Figure 7.79: Run Encoding: II. The encoder uses a 32-entry table J containing values that are denoted by rk. J is initialized to the 32 values 0, 0, 0, 0, 1, 1, 1, 1, 2, 2, 2, 2, 3, 3, 3, 3, 4, 4, 5, 5, 6, 6, 7, 7, 8, 9, 10, 11, 12, 13, 14, 15. For each value rk, we use the notation rm = 2rk. The 32 quantities rm are called code-order. The first four rms have values 20 = 1. The next four have values 21 = 2. The next four, 22 = 4, up to the last rm, whose value is 215 = 32768. The encoder executes the procedure of Figure 7.77 to determine the run length, which it stores in variable RUNlen. This variable is then encoded by breaking it up into chunks whose sizes are the values of consecutive rms. For example, if RUNlen is 6, it can be expressed 7.12 Adaptive Linear Prediction and Classification 547 in terms of the rms as 1 + 1 + 1 + 1 + 2, so it is equivalent to the first five rms. It is encoded by writing five bits of 1 on the compressed stream. Each of those bits is written by the statement AppendToBitStream(1,1) of Figure 7.78. Each time a 1 is written, the value of the corresponding rm is subtracted from RUNlen. If RUNlen is originally 6, it goes down to 5, 4, 3, 2, and 0. It may happen, of course, that the length RUNlen of a run is not equal to an integer number of rms. An example is a RUNlen of 7. This is encoded by writing five bits of 1, followed by a prefix bit, followed by the remainder of RUNlen (in our example, a 1), written on the compressed stream as an rk-bit number (the current rk in our example is 1). This last operation is performed by the procedure call Append- ToBitStream(RUNcnt,J[RUNindex]) of Figure 7.79. The prefix bit is 0 if the run is interrupted by a different pixel. It is 1 if the run is terminated by the end of an image row. The second main task of the encoder, encoding the interruption pixel, is similar to encoding the current pixel and is not discussed here. 7.12 Adaptive Linear Prediction and Classification ALPC (adaptive linear prediction and classification), is a lossless image compression algorithm described in [Motta et al. 00]. ALPC uses a linear predictor that is trained on a selected part of the causal neighborhood of the pixel being encoded. The use of linear prediction in an image compressor is somewhat unusual, since most lossless compressors rely on non-linear predictors to model the discontinuities often present in a natural image. In ALPC, the non-linearity of the input image is modeled by a classification step that is performed before the prediction. The classification selects a causal subset of samples from which the weights of the predictor are determined. Causality: A cause and effect relationship. The causality of two events describes to what extent one event is caused by the other. Causality implies a measure of predictability between the two events. In a linear predictor, a pixel value is represented by a weighted sum of its context pixels. ALPC uses a context formed by the six causal pixels depicted in Figure 7.80. The value of the pixel I(x, y) is predicted as ˆI(x, y) = round w0 Χ I(x, y - 2) + w1 Χ I(x - 1, y - 1) + w2 Χ I(x, y - 1) + w3 Χ I(x + 1, y - 1) + w4 Χ I(x - 2, y) + w5 Χ I(x - 1, y) . The shape of the context is fixed, while the weights w0 through w5 are adaptively determined for each pixel. Since the prediction is close to the value of the pixel, the prediction error E(x, y) = I(x, y)-ˆI(x, y) exhibits a peaked distribution centered around zero. As a result, encoding the prediction error is more effective than encoding the pixel itself. The image is encoded in raster scan order, from top to bottom and from left to right, so the decoder can produce an exact reconstruction of the context pixels and 548 7. Image Compression          Figure 7.80: Context Used in ALPC Linear Predictors. can mimic the encoder by computing ˆI(x, y) and adding it to E(x, y) to reconstruct the original value of I(x, y). The weights are computed on a pixel by pixel basis, in an attempt to model local characteristics of the image. The determination of the weights is based on the causal portion of a square window of radius Rp that is centered on the current pixel (Figure 7.81). This choice allows the use of backward prediction, a technique that does not require explicit transmission of the weights, because the decoder has enough information to determine them. The radius of the prediction window is a critical feature of ALPC. A small window will determine overspecialized predictors, while a large window will not be able to track the local variations in the image content. While it is possible to dynamically adapt Rp to the image content, in ALPC the radius of the window is fixed and is set equal to 10 pixels.          Figure 7.81: Prediction Window of Radius Rp Centered Around I(x, y). 7.13 Progressive Image Compression 549 Not all samples in the window are used to compute the predictors. Each pixel in the window is first classified in one of n classes, according to its context. The number of classes n is predetermined and a classification algorithm such as LBG [Linde, Buzo, and Gray 80] can be used to partition the samples in the window. The context of the current pixel is then classified as well and the contexts and the pixels belonging to its class are used to determine the predictor’s weights. While this classification method is not optimal, the selection of pixels having a context similar to the one being encoded is sufficient to improve the performance of the basic adaptive predictor. In [Motta et al. 00] the weights are determined by using a gradient descend optimization that minimizes the energy of the error on the samples in the class. If the number of samples is too small, a fixed predictor is used instead. The classified linear predictor is quite effective and the prediction error is distributed according to a skewed Laplacian-like distribution. In the experimental results reported in [Motta et al. 00], 95% of the error consistently falls into a very narrow range; we call these error values “typical.” Two arithmetic encoders are used to encode error magnitudes, depending on whether the prediction error is typical or not. The sign of the error is encoded separately, with a third arithmetic encoder. The modeling of the arithmetic encoders is based on error statistics collected in a causal window of radius Re, similar to the one used in the prediction. The use of multiple arithmetic encoders results in a small but consistent gain. 7.13 Progressive Image Compression Most modern image compression methods are either progressive or optionally so. Progressive compression is an attractive choice when compressed images are transmitted over a communications line and are decompressed and viewed in real time. When such an image is received and is decompressed, the decoder can very quickly display the entire image in a low-quality format, and improve the display quality as more and more of the image is being received and decompressed. A user watching the image developing on the screen can normally recognize most of the image features after only 5–10% of it has been decompressed. This should be compared to raster-scan image compression. When an image is raster scanned and compressed, a user normally cannot tell much about the image when only 5–10% of it has been decompressed and displayed. Images are supposed to be viewed by humans, which is why progressive compression makes sense even in cases where it is slower or less efficient than nonprogressive. Perhaps a good way to think of progressive image compression is to imagine that the encoder compresses the most important image information first, then compresses less important information and appends it to the compressed stream, and so on. This explains why all progressive image compression methods have a natural lossy option; simply stop compressing at a certain point. The user can control the amount of loss by means of a parameter that tells the encoder how soon to stop the progressive encoding process. The sooner encoding is stopped, the better the compression ratio and the higher the data loss. 550 7. Image Compression Another advantage of progressive compression becomes apparent when the compressed file has to be decompressed several times and displayed with different resolutions. The decoder can, in each case, stop the decompression when the image has reached the resolution of the particular output device used. Progressive image compression has already been mentioned, in connection with JPEG (page 522). JPEG uses the DCT to break the image up into its spatial frequency components, and it compresses the low-frequency components first. The decoder can therefore display these parts quickly, and it is these low-frequency parts that contain the principal image information. The high-frequency parts contain image details. Thus, JPEG encodes spatial frequency data progressively. It is useful to think of progressive decoding as the process of improving image features over time, and this can be achieved in three ways: 1. Encode spatial frequency data progressively. An observer watching such an image being decoded sees the image changing from blurred to sharp. Methods that work this way typically feature medium speed encoding and slow decoding. This type of progressive compression is sometimes called SNR progressive or quality progressive. 2. Start with a gray image and add colors or shades of gray to it. An observer watching such an image being decoded will see all the image details from the start, and will see them improve as more color is continuously added to them. Vector quantization methods (Section 7.19) use this kind of progressive compression. Such a method normally features slow encoding and fast decoding. 3. Encode the image in layers, where early layers consist of a few large low-resolution pixels, followed by later layers with smaller higher-resolution pixels. A person watching such an image being decoded will see more detail added to the image over time. Such a method thus adds detail (or resolution) to the image as it is being decompressed. This way of progressively encoding an image is called pyramid coding or hierarchical coding. Most progressive methods use this principle, so this section discusses general ideas for implementing pyramid coding. Figure 7.83 illustrates the three progressive methods mentioned here. It should be contrasted with Figure 7.82, which illustrates sequential decoding. Assuming that the image size is 2n Χ 2n = 4n pixels, the simplest method that comes to mind, when trying to implement progressive compression, is to calculate each pixel of layer i-1 as the average of a group of 2Χ2 pixels of layer i. Thus layer n is the entire image, layer n-1 contains 2n-1Χ2n-1 = 4n-1 large pixels of size 2Χ2, and so on, down to layer 1, with 4n-n = 1 large pixel, representing the entire image. If the image isn’t too large, all the layers can be saved in memory. The pixels are then written on the compressed stream in reverse order, starting with layer 1. The single pixel of layer 1 is the “parent” of the four pixels of layer 2, each of which is the parent of four pixels in layer 3, and so on. The total number of pixels in the pyramid is 33% more than the original number! 40 + 41 + · · · + 4n-1 + 4n = (4n+1 - 1)/3 . 4n(4/3) . 1.33Χ4n = 1.33(2nΧ2n). A simple way to bring the total number of pixels in the pyramid down to 4n is to include only three of the four pixels of a group in layer i, and to compute the value of 7.13 Progressive Image Compression 551 Figure 7.82: Sequential Decoding. the 4th pixel using the parent of the group (from the preceding layer, i-1) and its three siblings. Example: Figure 7.84c shows a 4Χ4 image that becomes the third layer in its progressive compression. Layer two is shown in Figure 7.84b, where, for example, pixel 81.25 is the average of the four pixels 90, 72, 140, and 23 of layer three. The single pixel of layer one is shown in Figure 7.84a. The compressed file should contain just the numbers 54.125, 32.5, 41.5, 61.25, 72, 23, 140, 33, 18, 21, 18, 32, 44, 70, 59, 16 (properly encoded, of course), from which all the missing pixel values can easily be determined. The missing pixel 81.25, e.g., can be calculated from (x + 32.5 + 41.5 + 61.25)/4 = 54.125. A small complication with this method is that averages of integers may be nonintegers. If we want our pixel values to remain integers we either have to lose precision or to keep using longer and longer integers. Assuming that pixels are represented by eight bits, adding four 8-bit integers produces a 10-bit integer. Dividing it by four, to create the average, reduces the sum back to an 8-bit integer, but some precision may be lost. If we don’t want to lose precision, we should represent our second-layer pixels as 10-bit numbers and our first-layer (single) pixel as a 12-bit number. Figure 7.84d,e,f shows the results of rounding off our pixel values and thus losing some image information. The content of the compressed file in this case should be 54, 33, 42, 61, 72, 23, 140, 33, 18, 21, 18, 32, 44, 70, 59, 16. The first missing pixel, 81, of layer three can be determined from the equation (x+33+ 42 + 61)/4 = 54, which yields the (slightly wrong) value 80. 552 7. Image Compression Figure 7.83: Progressive Decoding. 7.13 Progressive Image Compression 553 90 140 16 100 72 23 70 59 58 33 21 18 72 18 44 32 81.25 32.5 61.25 41.5 54.125 90 140 16 100 72 23 70 59 58 33 21 18 72 18 44 32 81 33 61 42 54 (a) (b) (c) (d) (e) (f) 90 140 16 min max max min max 100 72 23 70 59 58 33 21 18 72 18 44 32 140 21 16 72 140 (g) (h) (i) Figure 7.84: Progressive Image Compression.  Exercise 7.19: Show that the sum of four n-bit numbers is an (n + 2)-bit number. A better method is to let the parent of a group help in calculating the values of its four children. This can be done by calculating the differences between the parent and its children, and writing the differences (suitably coded) in layer i of the compressed stream. The decoder decodes the differences, then uses the parent from layer i - 1 to compute the values of the four pixels. Either Huffman or arithmetic coding can be used to encode the differences. If all the layers are calculated and saved in memory, then the distribution of difference values can be found and used to achieve the best statistical compression. If there is no room in memory for all the layers, a simple adaptive model can be implemented. It starts by assigning a count of 1 to every difference value (to avoid the zero-probability problem). When a particular difference is computed, it is assigned a probability and is encoded according to its count, and its count is then updated. It is a good idea to update the counts by incrementing them by a value greater than 1, since this way the original counts of 1 become insignificant very quickly. Some improvement can be achieved if the parent is used to help calculate the values of three child pixels, and then these three plus the parent are used to calculate the value of the fourth pixel of the group. If the four pixels of a group are a, b, c, and d, then their average is v = (a+b+c+d)/4. The average becomes part of layer i-1, and layer i need contain only the three differences k = a-b, l = b-c, and m = c-d. Once the decoder has read and decoded the three differences, it can use their values, together with the value of v from the previous layer, to compute the values of the four pixels of the group. Calculating v by a division by 4 still causes the loss of two bits, but this 2-bit quantity can be isolated before the division, and retained by encoding it separately, following the three differences. 554 7. Image Compression The improvements mentioned above are based on the well-known fact that small integers are easy to compress (page 46). The parent pixel of a group does not have to be its average. One alternative is to select the maximum (or the minimum) pixel of a group as the parent. This has the advantage that the parent is identical to one of the pixels in the group. The encoder has to encode just three pixels in each group, and the decoder decodes three pixels (or differences) and uses the parent as the fourth pixel, to complete the group. When encoding consecutive groups in a layer, the encoder should alternate between selecting the maximum and the minimum as parents, since always selecting the same creates progressive layers that are either too dark or too bright. Figure 7.84g,h,i shows the three layers in this case. The compressed file should contain the numbers 140, (0), 21, 72, 16, (3), 90, 72, 23, (3), 58, 33, 18, (0), 18, 32, 44, (3), 100, 70, 59, where the numbers in parentheses are two bits each. They tell where (in what quadrant) the parent from the previous layer should go. Notice that quadrant numbering is 0 1 3 2 . Selecting the median of a group is a little slower than selecting the maximum or the minimum, but it improves the appearance of the layers during progressive decompression. In general, the median of a sequence (a1, a2, . . . , an) is an element ai such that half the elements (or very close to half) are smaller than ai and the other half are bigger. If the four pixels of a group satisfy a < b < c < d, then either b or c can be considered the median pixel of the group. The main advantage of selecting the median as the group’s parent is that it tends to smooth large differences in pixel values that may occur because of one extreme pixel. In the group 1, 2, 3, 100, for example, selecting 2 or 3 as the parent is much more representative than selecting the average. Finding the median of four pixels requires a few comparisons, but calculating the average requires a division by 4 (or, alternatively, a right shift). Once the median has been selected and encoded as part of layer i-1, the remaining three pixels can be encoded in layer i by encoding their (three) differences, preceded by a 2-bit code telling which of the four is the parent. Another small advantage of using the median is that once the decoder reads this 2-bit code, it knows how many of the three pixels are smaller and how many are bigger than the median. If the code says, for example, that one pixel is smaller, and the other two are bigger than the median, and the decoder reads a pixel that’s smaller than the median, it knows that the next two pixels decoded will be bigger than the median. This knowledge changes the distribution of the differences, and it can be taken advantage of by using three count tables to estimate probabilities when the differences are encoded. One table is used when a pixel is encoded that the decoder will know is bigger than the median. Another table is used to encode pixels that the decoder will know are smaller than the median, and the third table is used for pixels where the decoder will not know in advance their relations to the median. This improves compression by a few percent and is another example of how adding more features to a compression method brings diminishing returns. Some of the important progressive image compression methods used in practice are described in the rest of this chapter. 7.13 Progressive Image Compression 555 7.13.1 Growth Geometry Coding The idea of growth geometry coding deserves its own section, since it combines image compression and progressive image transmission in an original and unusual way. This idea is due to Amalie J. Frank [Frank et al. 80]. The method is designed for progressive lossless compression of bi-level images. The idea is to start with some seed pixels and apply geometric rules to grow each seed pixel into a pattern of pixels. The encoder has the harder part of the job. It has to select the seed pixels, the growth rule for each seed, and the number of times the rule should be applied (the number of generations). Only this data is written on the compressed stream. Often, a group of seeds shares the same rule and the same number of generations. In such a case, the rule and number of generations are written once, following the coordinates of the seed pixels of the group. The decoder’s job is simple. It reads the first group of seeds, applies the rule once, reads the next group and adds it to the pattern so far, applies the rule again, and so on. Compression is achieved if the number of seeds is small compared to the total image size, i.e., if each seed pixel is used to generate, on average, many image pixels. (g) (h) (i) (j) (k) (m) (a) (b) (c) (d) (e) (f) 2 2 2 2 1 2 2 2 2 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 2 3 3 4 1 4 2 2 2 1 2 2 3 3 3 3 3 3 3 3 3 3 3 3 2 2 3 2 1 S S S S S S S’ 2 2 2 2 1 2 2 2 2 3 3 3 3 2 1 2 3 3 Figure 7.85: Six Growth Rules and Patterns. Figure 7.85a–f shows six simple growth rules. Each adds some immediate neighbor pixels to the seed S, and the same rule is applied to these neighbors in the next generation (they become secondary seed pixels). Since a pixel can have up to eight immediate neighbors, there can be 256 such rules. Figure 7.85g–m shows the results of applying these rules twice (i.e., for two generations) to a single seed pixel. The pixels are numbered one plus the generation number. The growth rules can be as complex as necessary. For example, Figure 7.85m assumes that only pixel S becomes a secondary seed. The resulting pattern of pixels may be solid or may have “holes” in it, as in Figure 7.85k. Figure 7.86 shows an example of how a simple growth rule (itself shown in Figure 7.86a) is used to grow the disconnected pattern of Figure 7.86b. The seed pixels are shown in Figure 7.86c. They are numbered 1–4 (one more than the number of generations). The decoder first inputs the six pixels marked 4 (Figure 7.86d). The growth rule is applied to them, producing the pattern of Figure 7.86e. The single pixel marked 3 is then input, and the rule applied again, to pixels 3 and 4, to form the pattern of Figure 7.86f. The four pixels marked 2 are input, and the rule applied again, to form 556 7. Image Compression (a) (b) (c) (d) (e) (f) (g) 2 2 4 4 4 3 4 4 2 4 1 1 1 1 1 1 1 1 1 1 2 2 2 4 4 4 3 4 4 2 4 1 1 1 1 1 1 1 1 1 1 2 2 2 4 4 4 3 4 4 2 4 2 4 4 4 3 4 4 4 4 4 4 4 4 4 S Figure 7.86: Growing a Pattern. the pattern of Figure 7.86g. Finally, the ten pixels marked 1 are read, to complete the image. The number of generations for these pixels is zero; they are not used to grow any other pixels, and so they do not contribute to the image compression. Notice that the pixels marked 4 go through three generations. In this example, the encoder wrote the seed pixels on the compressed stream in order of decreasing number of generations. In principle, it is possible to complicate matters as much as desired. It is possible to use different growth rules for different groups of pixels (in such a case, the growth rules must tell unambiguously who the parent of a pixel P is, so that the decoder will know what growth rule to use for P), to stop growth if a pixel comes to within a specified distance of other pixels, to change the growth rule when a new pixel bumps into another pixel, to reverse direction in such a case and start erasing pixels, or to use any other algorithm. In practice, however, complex rules require more bits to encode them, which is why simple rules may lead to better compression.  Exercise 7.20: Simple growth rules have another advantage. What is it? Figure 7.87 illustrates a simple approach to the design of a recursive encoder that identifies the seed pixels in layers. We assume that the only growth rule used is the one shown in Figure 7.87a. All the black (foreground) pixels are initially marked 1, and the white (background) pixels (not shown in the figure) are marked 0. The encoder scans the image and marks by 2 each 1-pixel that can grow. In order for a 1-pixel to grow it must be surrounded by 1-pixels on all four sides. The next layer consists of the resulting 2-pixels, shown in Figure 7.87c. The encoder next looks for 1-pixels that do not have at least one flanking 2-pixel. These are seed pixels. There are 10 of them, marked with a 7.14 JBIG 557 (a) (b) (c) (d) (e) (f) (g) 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 2 2 2 2 2 2 2 2 2 2 1 2 2 1 1 1 2 1 1 1 2 2 2 1 1 1 2 2 2 1 1 1 1 2 2 2 2 2 1 2 2 1 1 1 1 2 2 2 1 1 2 1 1 1 2 1 2 2 2 2 2 2 2 4 4 4 3 4 4 2 4 1 1 1 1 1 1 1 1 1 1 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 3 3 3 3 3 3 3 3 3 3 3 3 3 3 2 3 3 3 2 4 4 4 3 3 3 3 3 3 3 3 3 3 4 4 3 4 4 4 4 4 4 4 S Figure 7.87: Recursive Encoding of a Pattern. gray background in Figure 7.87c. The same process is applied to the layer of 2-pixels. The encoder scans the image and marks by 3 each 2-pixel that can grow. The next (third) layer consists of the resulting 3-pixels, shown in Figure 7.87d. The encoder next looks for 2-pixels that do not have at least one flanking 3-pixel. These are seed pixels. There are four of them, marked with a gray background in Figure 7.87d. Figure 7.87e shows the fourth layer. It consists of six 4-pixels and one seed 3-pixel (marked in gray). Figure 7.87f shows the six seed 4-pixels, and the final seed pixels are shown in Figure 7.87g. This basic algorithm can be extended in a number of ways, most notably by using several growth rules for each layer, instead of just one. 7.14 JBIG No compression method can efficiently compress every type of data. This is why new special-purpose methods are being developed all the time. JBIG [JBIG 03] is an example of a special-purpose method. It has been developed specifically for progressive compression of bi-level images. Such images, also called monochromatic or black and white, are common in applications where drawings (technical or artistic), with or without text, need to be saved in a database and retrieved. It is customary to use the terms “foreground” and “background” instead of black and white, respectively. The term “progressive compression” means that the image is saved in several “layers” in the compressed stream, at higher and higher resolutions. When such an image 558 7. Image Compression is decompressed and viewed, the viewer first sees an imprecise, rough, image (the first layer) followed by improved versions of it (later layers). This way, if the image is the wrong one, it can be rejected at an early stage, without having to retrieve and decompress all of it. Section 7.13 shows how each high-resolution layer uses information from the preceding lower-resolution layer, so there is no duplication of data. This feature is supported by JBIG, where it is an option called deterministic prediction. Even though JBIG was designed for bi-level images, where each pixel is one bit, it is possible to apply it to grayscale images by separating the bitplanes and compressing each individually, as if it were a bi-level image. RGC (reflected Gray code) should be used, instead of the standard binary code, as discussed in Section 7.4.1. The name JBIG stands for Joint Bi-Level Image Processing Group. This is a group of experts from several international organizations, formed in 1988 to recommend such a standard. The official name of the JBIG method is ITU-T recommendation T.82. ITU is the International Telecommunications Union (part of the United Nations). The ITU-T is the telecommunication standardization sector of the ITU. JBIG uses multiple arithmetic coding to compress the image, and this part of JBIG, which is discussed below, is separate from the progressive compression part discussed in Section 7.14.1. An important feature of the definition of JBIG is that the operation of the encoder is not defined in detail. The JBIG standard discusses the details of the decoder and the format of the compressed file. It is implied that any encoder that generates a JBIG file is a valid JBIG encoder. The JBIG2 method of Section 7.15 adopts the same approach, because this allows implementers to come up with sophisticated encoders that analyze the original image in ways that could not have been envisioned at the time the standard was developed. One feature of arithmetic coding is that it is easy to separate the statistical model (the table with frequencies and probabilities) from the encoding and decoding operations. It is easy to encode, for example, the first half of a data stream using one model and the second half using another model. This is called multiple arithmetic coding, and it is especially useful in encoding images, since it takes advantage of any local structures and interrelationships that might exist in the image. JBIG uses multiple arithmetic coding with many models, each a two-entry table that gives the probabilities of a white and a black pixel. There are between 1,024 and 4,096 such models, depending on the resolution of the image that’s being compressed. A bi-level image is made up of foreground (black) and background (white) dots called pixels. The simplest way to compress such an image with arithmetic coding is to count the frequency of black and white pixels, and compute their probabilities. In practice, however, the probabilities vary from region to region in the image, and this fact can be used to produce better compression. Certain regions, such as the margins of a page, may be completely white, while other regions, such as the center of a complex diagram, or a large, thick rule, may be predominantly or completely black. Thus, pixel distributions in adjacent regions in an image can vary wildly. Consider an image where 25% of the pixels are black. The entropy of this image is -0.25 log2 0.25 - 0.75 log2 0.75 . 0.8113. The best that we can hope for is to represent each pixel with 0.81 bits instead of the original 1 bit: an 81% compression ratio (or 0.81 bpp). Now assume that we discover that 80% of the image is predominantly white, with just 10% black pixels, and the remaining 20% have 85% black pixels. The entropies of 7.14 JBIG 559 these parts are -0.1 log2 0.1 - 0.9 log2 0.9 . 0.47 and -0.85 log2 0.85 - 0.15 log2 0.15 . 0.61, so if we encode each part separately, we can have 0.47 bpp 80% of the time and 0.61 bpp the remaining 20%. On average this results in 0.498 bpp, or a compression ratio of about 50%; much better than 81%! We assume that a white pixel is represented by a 0 and a black one by a 1. In practice, we don’t know in advance how many black and white pixels exist in each part of the image, so the JBIG encoder stops at every pixel and examines a template made of the 10 neighboring pixels marked (in Figures 7.88 and 7.89) “X” and “A” above it and to its left (those that have already been input; the ones below and to the right are still unknown). It interprets the values of these 10 pixels as a 10-bit integer which is then used as a pointer to a statistical model, which, in turn, is used to encode the current pixel (marked by a “?”). There are 210 = 1,024 10-bit integers, so there should be 1,024 models. Each model is a small table consisting of the probabilities of black and white pixels (just one probability needs be saved in the table, since the probabilities add up to 1). Figure 7.88a,b shows the two templates used for the lowest-resolution layer. The encoder decides whether to use the three-line or the two-line template and sets parameter LRLTWO in the compressed file to 0 or 1, respectively, to indicate this choice to the decoder. (The two-line template results in somewhat faster execution, while the three-line template produces slightly better compression.) Figure 7.88c shows the 10-bit template 0001001101, which becomes the pointer 77. The pointer shown in Figure 7.88d for pixel Y is 0000100101 or 37. Whenever any of the pixels used by the templates lies outside the image, the JBIG edge convention mentioned earlier should be used. Figure 7.89 shows the four templates used for all the other layers. These templates reflect the fact that when any layer, except the first one, is decoded the low-resolution pixels from the preceding layer are known to the decoder. Each of the four templates is used to encode (and later decode) one of the four high-resolution pixels of a group. The context (pointer) generated in these cases consists of 12 bits, 10 taken from the template’s pixels and two generated to indicate which of the four pixels in the group is being processed. The number of statistical models should therefore be 212 = 4096. The two bits indicating the position of a high-resolution pixel in its group are 00 for the top-left pixel (Figure 7.89a), 01 for the top-right (Figure 7.89b), 10 for the bottom-left (Figure 7.89c), and 11 for the bottom-right pixel (Figure 7.89d). The use of these templates implies that the JBIG probability model is a 10th-order or a 12th-order Markov model. The template pixels labeled A are called adaptive pixels (AT). The encoder is allowed to use as AT a pixel other than the one specified in the template, and it uses two parameters Tx and Ty (one byte each) in each layer to indicate to the decoder the actual location of the AT in that layer. (The AT is not allowed to overlap one of the X pixels in the template.) A sophisticated JBIG encoder may notice, for example, that the image uses halftone patterns where a black pixel normally has another black pixel three rows above it. In such a case the encoder may decide to use the pixel three rows above ? as the AT. Figure 7.90 shows the image area where AT may reside. Parameter Mx can vary in the range [0, 127], while My can take the values 0 through 255. 560 7. Image Compression O O O O O O O A O O ? O O O O O A O O O O ? (a) (b) 0 0 0 1 0 0 1 1 0 1 X 0 0 0 0 1 0 0 1 0 1 Y (c) (d) Figure 7.88: Templates for Lowest-Resolution Layer. (a) (b) (c) (d) A X X X X X ? A X X X X X ? A X X X X X ? A X X X X X ? X X X X X X X X X X X X X X X X Figure 7.89: Templates for Other Layers. 7.14 JBIG 561 Mx,My · · · 2,My 1,My 0,My -1,My -2,My · · ·-Mx,My ... ... ... Mx, 1 · · · 2, 1 1, 1 0, 1 -1, 1 -2, 1 · · · -Mx, 1 Mx, 0 · · · 2, 0 1, 0 ? Figure 7.90: Coordinates and Allowed Positions of AT Pixels. 7.14.1 Progressive Compression One advantage of the JBIG method is its ability to generate low-resolution versions (layers) of the image in the compressed stream. The decoder decompresses these layers progressively, from the lowest to the highest resolution. We first look at the order of the layers. The encoder is given the entire image (the highest-resolution layer), so it is natural for it to construct the layers from high to low resolution and write them on the compressed file in this order. The decoder, on the other hand, has to start by decompressing and displaying the lowest-resolution layer, so it is easiest for it to read this layer first. As a result, either the encoder or the decoder should use buffering to reverse the order of the layers. If fast decoding is important, the encoder should use buffers to accumulate all the layers and then write them on the compressed file from low to high resolutions. The decoder then reads the layers in the right order. In cases where fast encoding is important (such as an archive that’s being updated often but is rarely decompressed and used), the encoder should write the layers in the order in which they are generated (from high to low) and the decoder should use buffers. The JBIG standard allows either method, and the encoder has to set a bit denoted by HITOLO in the compressed file to either zero (layers are in low to high order) or one (the opposite case). It is implied that the encoder decides what the resolution of the lowest layer should be (it doesn’t have to be a single pixel). This decision may be based on user input or on information built in the encoder about the specific needs of a particular environment. Progressive compression in JBIG also involves the concept of stripes. A stripe is a narrow horizontal band consisting of L scan lines of the image, where L is a usercontrolled JBIG parameter. As an example, if L is chosen such that the height of a stripe is 8 mm (about 0.3 inches), then there will be about 36 stripes in an 11.5-inchhigh image. The encoder writes the stripes on the compressed file in one of two ways, setting parameter SET to indicate this to the decoder. Either all the stripes of a layer are written on the compressed file consecutively, followed by the stripes of the next layer, and so on (SET=0), or the top stripes of all the layers are first written, followed by the second stripes of all the layers, and so on (SET=1). If the encoder sets SET=0, then 562 7. Image Compression the decoder will decompress the image progressively, layer by layer, and in each layer stripe by stripe. Figure 7.91 illustrates the case of an image divided into four stripes and encoded in three layers (from low-resolution 150 dpi to high-resolution 600 dpi). Table 7.92 lists the order in which the stripes are output for the four possible values of parameters HITOLO and SET. 0 1 2 3 4 5 6 7 8 9 10 11 layer 1 150 dpi layer 2 300 dpi layer 3 600 dpi Figure 7.91: Four Stripes and Three Layers in a JBIG Image. HITOLO SEQ Order 0 0 0,1,2,3,4,5,6,7,8,9,10,11 0 1 0,4,8,1,5,9,2,6,10,3,7,11 1 0 8,9,10,11,4,5,6,7,0,1,2,3 1 1 8,4,0,9,5,1,10,6,2,11,7,3 Table 7.92: The Four Possible Orders of Layers. The basic idea of progressive compression is to group four high-resolution pixels into one low-resolution pixel, a process called downsampling. The only problem is to determine the value (black or white) of that pixel. If all four original pixels have the same value, or even if three are identical, the solution is obvious (follow the majority). When two pixels are black and the other two are white, we can try one of the following solutions: 1. Create a low-resolution pixel that’s always black (or always white). This is a bad solution, since it may eliminate important details of the image, making it impractical or even impossible for an observer to evaluate the image by viewing the low-resolution layer. 2. Assign a random value to the new low-resolution pixel. This solution is also bad, since it may add too much noise to the image. 3. Give the low-resolution pixel the color of the top-left of the four high-resolution pixels. This prefers the top row and the left column, and has the drawback that thin lines can sometimes be completely missing from the low-resolution layer. Also, if the high-resolution layer uses halftoning to simulate grayscales, the halftone patterns may be corrupted. 4. Assign a value to the low-resolution pixel that depends on the four high-resolution pixels and on some of their nearest neighbors. This solution is used by JBIG. If most of 7.14 JBIG 563 the near neighbors are white, a white low-resolution pixel is created; otherwise, a black low-resolution pixel is created. Figure 7.93a shows the 12 high-resolution neighboring pixels and the three low-resolution ones that are used in making this decision. A, B, and C are three low-resolution pixels whose values have already been determined. Pixels d, e, f, g, and, j are on top or to the left of the current group of four pixels. Pixel “?” is the low-resolution pixel whose value needs to be determined. e f g h i j A C B ? d k l 2 1 2 4 2 1 -1 ? 1 2 1 -3 -3 (a) (b) Figure 7.93: HiRes and LoRes Pixels. Figure 7.93b shows the weights assigned to the various pixels involved in the determination of pixel “?”. The weighted sum of the pixels can also be written as 4h + 2(e + g + i + k) + (d + f + j + l) - 3(B + C) - A = 4h + 2(i + k) + l + (d - A) + 2(g - C) + (j - C) + 2(e - B) + (f - B). (7.26) The second line of Equation (7.26) shows how the value of pixel “?” depends on differences such as d-A (a difference of a high-resolution pixel and the adjacent low-resolution pixel). Assuming equal probabilities for black and white pixels, the first line of Equation (7.26) can have values between zero (when all 12 pixels are white) and 9 (when all are black), so the rule is to assign pixel “?” the value 1 (black) if expression (7.26) is greater than 4.5 (if it is 5 or more), and the value 0 (white) otherwise (if it is 4 or less). Thus, this expression acts as a filter that preserves the density of the high-resolution pixels in the low-resolution image.  Exercise 7.21: We normally represent a black (or foreground) pixel with a binary 1, and a white (or background) pixel with a binary 0. How will the resolution reduction method above change if we use 0 to represent black and 1 to represent white? Since the image is divided into groups of 4Χ4 pixels, it should have an even number of rows and columns. The JBIG edge convention says that an image can be extended if and when necessary by adding columns of 0 pixels at the left and right, rows of 0 pixels at the top, and by replicating the bottom row as many times as necessary. The JBIG method of resolution reduction has been carefully designed and extensively tested. It is known to produce excellent results for text, drawings, halftoned grayscales, as well as for other types of images. 564 7. Image Compression JBIG also includes exceptions to the above rule in order to preserve edges (132 exceptions), preserve vertical and horizontal lines (420 exceptions), periodic patterns in the image (10 exceptions, for better performance in regions of transition to and from periodic patterns), and dither patterns in the image (12 exceptions that help preserve very low density or very high density dithering, i.e., isolated background or foreground pixels). The last two groups are used to preserve certain shading patterns and also patterns generated by halftoning. Each exception is a 12-bit number. Table 7.94 lists some of the exception patterns. The pattern of Figure 7.95b was developed in order to preserve thick horizontal, or near-horizontal, lines. A two-pixel-wide high-resolution horizontal line, e.g., will result in a low-resolution line whose width alternates between one and two low-resolution pixels. When the upper row of the high-resolution line is scanned, it will result in a row of low-resolution pixels because of Figure 7.95a. When the next high-resolution row is scanned, the two rows of the thick line will generate alternating low-resolution pixels because pattern 7.95b requires a zero in the bottom-left low-resolution pixel in order to complement the bottom-right low-resolution pixel. Pattern 7.95b leaves two pixels unspecified, so it counts for four patterns. Its reflection is also used, to preserve thick vertical lines, so it counts for eight of the 420 line-preservation exceptions. Figure 7.95c complements a low-resolution pixel in cases where there is one black high-resolution pixel among 11 white ones. This is common when a grayscale image is converted to black and white by halftoning. This pattern is one of the 12 dither preservation exceptions. The JBIG method for reducing resolution seems complicated—especially since it includes so many exceptions—but is actually simple to implement and also executes very fast. The decision whether to paint the current low-resolution pixel black or white depends on 12 of its neighboring pixels, nine high-resolution pixels, and three lowresolution pixels. Their values are combined into a 12-bit number, which is used as a pointer to a 4,096-entry table. Each entry in the table is one bit, which becomes the value of the current low-resolution pixel. This way, all the exceptions are already included in the table and don’t require any special treatment by the program. It has been mentioned in Section 7.13 that compression is improved if the encoder writes on the compressed stream only three of the four pixels of a group in layer i. This is enough because the decoder can compute the value of the 4th pixel from its three siblings and the parent of the group (the parent is from the preceding layer, i - 1). Equation (7.26) shows that in JBIG, a low-resolution pixel is not calculated as a simple average of its four high-resolution “children,” so it may not be possible for the JBIG decoder to calculate the value of, say, high-resolution pixel k from the values of its three siblings h, i, and l and their parent. The many exceptions used by JBIG also complicate this calculation. As a result, JBIG uses special tables to tell both encoder and decoder which high-resolution pixels can be inferred from their siblings and parents. Such pixels are not compressed by the encoder, but all the other high-resolution pixels of the layer are. This method is referred to as deterministic prediction (DP), and is a JBIG option. If the encoder decides to use it, it sets parameter DPON to 1. Figure 7.96 shows the pixel numbering used for deterministic prediction. Before it compresses high-resolution pixel 8, the encoder combines the values of pixels 0–7 to 7.14 JBIG 565 The 10 periodic pattern preservation exceptions are (in hex) 5c7 36d d55 b55 caa aaa c92 692 a38 638. The 12 dither pattern preservation exceptions are fef fd7 f7d f7a 145 142 ebd eba 085 082 028 010. The 132 edge-preservation exception patterns are a0f 60f 40f 20f e07 c07 a07 807 607 407 207 007 a27 627 427 227 e1f 61f e17 c17 a17 617 e0f c0f 847 647 447 247 e37 637 e2f c2f a2f 62f e27 c27 24b e49 c49 a49 849 649 449 249 049 e47 c47 a47 e4d c4d a4d 84d 64d 44d 24d e4b c4b a4b 64b 44b e69 c69 a69 669 469 269 e5b 65b e59 c59 a59 659 ac9 6c9 4c9 2c9 ab6 8b6 e87 c87 a87 687 487 287 507 307 cf8 8f8 ed9 6d9 ecb ccb acb 6cb ec9 cc9 949 749 549 349 b36 336 334 f07 d07 b07 907 707 3b6 bb4 3b4 bb2 3a6 b96 396 d78 578 f49 d49 b49 ff8 df8 5f8 df0 5f0 5e8 dd8 5d8 5d0 db8 fb6 bb6 where the 12 bits of each pattern are numbered as in the diagram. 7 6 5 4 3 2 8 1 0 11 10 9 ? Table 7.94: Some JBIG Exception Patterns. Figure 7.95 shows three typical exception patterns. A pattern of six zeros and three ones, as in Figure 7.95a, is an example exception, introduced to preserve horizontal lines. It means that the low-resolution pixel marked “C” should be complemented (assigned the opposite of its normal value). The normal value of this pixel depends, of course, on the three low-resolution pixels above it and to the left. Since these three pixels can have eight different values, this pattern covers eight exceptions. Reflecting 7.95a about its main diagonal produces a pattern (actually, eight patterns) that’s natural to use as an exception to preserve vertical lines in the original image. Thus, Figure 7.95a corresponds to 16 of the 420 line-preservation exceptions. 0 0 0 0 0 1 0 1 1 C 1 1 1 1 0 1 0 0 C 1 0 0 0 0 1 0 0 0 0 0 C 0 0 0 (a) (b) (c) Figure 7.95: Some JBIG Exception Patterns. 566 7. Image Compression 5 6 7 8 9 10 4 11 12 0 1 2 3 Figure 7.96: Pixel Numbering for Deterministic Prediction. form an 8-bit pointer to the first DP table. If the table entry is 2, then pixel 8 should be compressed, since the decoder cannot infer it from its neighbors. If the table entry is 0 or 1, the encoder ignores pixel 8. The decoder prepares the same pointer (since pixels 0–7 are known when pixel 8 should be decoded). If the table entry is 0 or 1, the decoder assigns this value to pixel 8; otherwise (if the table entry is 2), the next pixel value is decoded from the compressed file and is assigned to pixel 8. Notice that of the 256 entries of Table 7.97 only 20 are not 2. Pixel 9 is handled similarly, except that the pointer used consists of pixels 0–8, and it points to an entry in the second DP table. This table, shown in 7.98, consists of 512 entries, of which 108 are not 2. The JBIG standard also specifies the third DP table, for pixel 10 (1024 entries, of which 526 are not 2), and the fourth table, for pixel 11 (2048 entries, of which 1044 are not 2). The total size of the four tables is 3840 entries, of which 1698 (or about 44%) actually predict pixel values. Pointer Value 0–63 02222222 22222222 22222222 22222222 02222222 22222222 22222222 22222222 64–127 02222222 22222222 22222222 22222222 00222222 22222222 22222222 22222222 128–191 02222222 22222222 00222222 22222222 02020222 22222222 02022222 22222222 192–255 00222222 22222222 22222222 22222221 02020022 22222222 22222222 22222222 Table 7.97: Predicting Pixel 8. Pointer Value 0–63 22222222 22222222 22222222 22000000 02222222 22222222 00222222 22111111 64–127 22222222 22222222 22222222 21111111 02222222 22111111 22222222 22112221 128–191 02222222 22222222 02222222 22222222 00222222 22222200 20222222 22222222 192–255 02222222 22111111 22222222 22222102 11222222 22222212 22220022 22222212 256–319 20222222 22222222 00220222 22222222 20000222 22222222 00000022 22222221 320–383 20222222 22222222 11222222 22222221 22222222 22222221 22221122 22222221 384–447 20020022 22222222 22000022 22222222 20202002 22222222 20220002 22222222 448–511 22000022 22222222 00220022 22222221 21212202 22222222 22220002 22222222 Table 7.98: Predicting Pixel 9. 7.15 JBIG2 567 7.15 JBIG2 They did it again! The Joint Bi-Level Image Processing Group has produced another standard for the compression of bi-level images. The new standard is called JBIG2, implying that the old standard should be called JBIG1. JBIG2 was developed in the late 1990s and was approved by the ISO and the ITU-T in 1999. Current information about JBIG2 can be found at [JBIG2 03] and [JBIG2 06]. The JBIG2 standard offers: 1. Large increases in compression performance (typically 3–5 times better than Group 4/MMR, and 2–4 times better than JBIG1). 2. Special compression methods for text, halftones, and other bi-level image parts. 3. Lossy and lossless compression. 4. Two modes of progressive compression. Mode 1 is quality-progressive compression, where the decoded image progresses from low to high quality. Mode 2 is contentprogressive coding, where important image parts (such as text) are decoded first, followed by less-important parts (such as halftone patterns). 5. Multipage document compression. 6. Flexible format, designed for easy embedding in other image file formats. 7. Fast decompression: In some coding modes, images can be decompressed at over 250 million pixels/second in software. The JBIG2 standard describes the principles of the compression and the format of the compressed file. It does not get into the operation of the encoder. Any encoder that produces a JBIG2 compressed file is a valid JBIG2 encoder. It is hoped that this will encourage software developers to implement sophisticated JBIG2 encoders. The JBIG2 decoder reads the compressed file, which contains dictionaries and image information, and decompresses it into the page buffer. The image is then displayed or printed from the page buffer. Several auxiliary buffers may also be used by the decoder. A document to be compressed by JBIG2 may consist of more than one page. The main feature of JBIG2 is that it distinguishes between text, halftone images, and everything else on the page. The JBIG2 encoder should somehow scan the page before doing any encoding, and identify regions of three types: 1. Text regions. These contain text, normally arranged in rows. The text does not have to be in any specific language, and may consist of unknown symbols, dingbats, musical notation, or hieroglyphs. The encoder treats each symbol as a rectangular bitmap, which it places in a dictionary. The dictionary is compressed by either arithmetic coding (specifically, a version of arithmetic coding called the MQ-coder, [Pennebaker and Mitchell 88a,b]) or MMR (Section 5.7.2) and written, as a segment, on the compressed file. Similarly, the text region itself is encoded and written on the compressed file as another segment. To encode a text region the encoder prepares the relative coordinates of each symbol (the coordinates relative to the preceding symbol), and a pointer to the symbol’s bitmap in the dictionary. (It is also possible to have pointers to several bitmaps and specify the symbol as an aggregation of those bitmaps, such as logical AND, OR, or XOR.) This data also is encoded before being written on the compressed file. Compression is achieved if the same symbol occurs several times. Notice that the symbol may occur in different text regions and even on different pages. The encoder should make sure that the dictionary containing this symbol’s bitmap will be retained by 568 7. Image Compression the decoder for as long as necessary. Lossy compression is achieved if the same bitmap is used for symbols that are slightly different. 2. Halftone regions. A bi-level image may contain a grayscale image done in halftone [Salomon 99]. The encoder scans such a region cell by cell (halftone cells, also called patterns, are typically 3Χ3 or 4Χ4 pixels). A halftone dictionary is created by collecting all the different cells and converting each to an integer (typically nine bits or 16 bits). The dictionary is compressed and is written on the output file. Each cell in the halftone region is then replaced by a pointer to the dictionary. The same dictionary can be used to decode several halftone dictionaries, perhaps located on different pages of the document. The encoder should label each dictionary so the decoder would know how long to retain it. 3. Generic regions. This is any region not identified by the encoder as text or halftone. Such a region may contain a large character, line art, mathematics, or even noise (specks and dirt). A generic region is compressed with either arithmetic coding or MMR. When the former is used, the probability of each pixel is determined by its context (i.e., by several pixels located above it and to its left that have already been encoded). A page may contain any number of regions, and they may overlap. The regions are determined by the encoder, whose operation is not specified by the JBIG2 standard. As a result, a simple JBIG2 encoder may encode any page as one large generic region, while a sophisticated encoder may spend time analyzing the contents of a page and identifying several regions, thereby leading to better compression. Figure 7.99 is an example of a page with four text regions, one halftone region, and two generic region (a sheared face and fingerprints). The fact that the JBIG2 standard does not specify the encoder is also used to generate lossy compression. A sophisticated encoder may decide that dropping certain pixels would not deteriorate the appearance of the image significantly. The encoder therefore drops those bits and creates a small compressed file. The decoder operates as usual and does not know whether the image that’s being decoded is lossy or not. Another approach to lossy compression is for the encoder to identify symbols that are very similar and replace them by pointers to the same bitmap in the symbol dictionary (this approach is risky, since symbols that differ by just a few pixels may be completely different to a human reader, think of “e” and “c” or “i” and “ύ”). An encoder may also be able to identify specks and dirt in the document and optionally ignore them. JBIG2 introduces the concept of region refinement. The compressed file may include instructions directing the decoder to decode a region A from the compressed file into an auxiliary buffer. That buffer may later be used to refine the decoding of another region B. When B is found in the compressed file and is decoded into the page buffer, each pixel written in the page buffer is determined by pixels decoded from the compressed file and by pixels in the auxiliary buffer. One example of region refinement is a document where certain text regions contain the words “Top Secret” in large, gray type, centered as a background. A simple JBIG2 encoder may consider each of the pixels constituting “Top Secret” part of some symbol. A sophisticated encoder may identify the common background, treat it as a generic region, and compress it as a region refinement, to be applied to certain text regions by the decoder. Another example of region refinement is a document where certain halftone regions have dark background. A sophisticated encoder may identify the dark background in those regions, treat it as a generic region, 7.15 JBIG2 569 1. Text regions. These contain text, normally arranged in rows. The text does not have to be in any specific language, and may consist of unknown symbols, dingbats, music notes, or hieroglyphs. The encoder treats each symbol as a rectangular bitmap which it places in a dictionary. The dictionary is compressed by either arithmetic coding or MMR and written, as a segment, on the compressed file. Similarly, the text region itself is encoded and written on the compressed file as another segment. To encode a text region the encoder prepares the relative coordinates of each symbol (the coordinates relative to the preceding symbol), and a pointer to the symbol’s bitmap in the dictionary. (It is also possible to have pointers to several bitmaps and specify the symbol as an aggregation of those bitmaps, such as logical AND, OR, or XOR.) This data also is encoded before being written on the compressed file. Compression is achieved if the same symbol occurs several times. Notice that the symbol may occur in different text regions. Lena in halftone (2 by 2 inches) 2. Halftone regions. A bi-level image may contain a grayscale image done in halftone. The encoder scans such a region cell by cell (halftone cells, also called patterns, are typically... A halftone dictionary is created by collecting all the different cells, and converting each to an integer (typically 9 bits or 16 bits). Example of shearing Figure 7.99: Typical JBIG2 Regions. and place instructions in the compressed file telling the decoder to use this region to refine certain halftone regions. Thus, a simple JBIG2 encoder may never use region refinement, but a sophisticated one may use this concept to add special effects to some regions. The decoder starts by initializing the page buffer to a certain value, 0 or 1, according to a code read from the compressed file. It then inputs the rest of the file segment by segment and executes each segment by a different procedure. There are seven main procedures. 1. The procedure to decode segment headers. Each segment starts with a header that includes, among other data and parameters, the segment’s type, the destination of the decoded output from the segment, and which other segments have to be used in decoding this segment. 570 7. Image Compression 2. A procedure to decode a generic region. This is invoked when the decoder finds a segment describing such a region. The segment is compressed with either arithmetic coding or MMR, and the procedure decompresses it (pixel by pixel in the former case and runs of pixels in the latter). In the case of arithmetic coding, previously decoded pixels are used to form a prediction context. Once a pixel is decoded, the procedure does not simply store it in the page buffer, but combines it with the pixel already in the page buffer according to a logical operation (AND, OR, XOR, or XNOR) specified in the segment. 3. A procedure to decode a generic refinement region. This is similar to the above except that it modifies an auxiliary buffer instead of the page buffer. 4. A procedure to decode a symbol dictionary. This is invoked when the decoder finds a segment containing such a dictionary. The dictionary is decompressed and is stored as a list of symbols. Each symbol is a bitmap that is either explicitly specified in the dictionary or is specified as a refinement (i.e., a modification) of a known symbol (a preceding symbol from this dictionary or a symbol from another existing dictionary) or is specified as an aggregate (a logical combination) of several known symbols). 5. A procedure to decode a symbol region. This is invoked when the decoder finds a segment describing such a region. The segment is decompressed and it yields triplets. The triplet for a symbol contains the coordinates of the symbol relative to the preceding symbol and a pointer (index) to the symbol in the symbol dictionary. Since the decoder may keep several symbol dictionaries at any time, the segment should indicate which dictionary is to be used. The symbol’s bitmap is brought from the dictionary, and the pixels are combined with the pixels in the page buffer according to the logical operation specified by the segment. 6. A procedure to decode a halftone dictionary. This is invoked when the decoder finds a segment containing such a dictionary. The dictionary is decompressed and is stored as a list of halftone patterns (fixed-size bitmaps). 7. A procedure to decode a halftone region. This is invoked when the decoder finds a segment describing such a region. The segment is decompressed into a set of pointers (indexes) to the patterns in the halftone dictionary. Some of these procedures are described here in more detail. 7.15.1 Generic Region Decoding This procedure reads several parameters from the compressed file (Table 7.100, where “I” stands for integer and “B” stands for bitmap), the first of which, MMR, specifies the compression method used in the segment about to be decoded. Either arithmetic coding or MMR can be used. The former is similar to arithmetic coding in JBIG1 and is described below. The latter is used as in fax compression (Section 5.7). If arithmetic coding is used, pixels are decoded one by one and are placed row by row in the generic region being decoded (part of the page buffer), whose width and height are given by parameters GBW and GBH, respectively. They are not simply stored there but are logically combined with the existing background pixels. Recall that the decompression process in arithmetic coding requires knowledge of the probabilities of the items being decompressed (decoded). These probabilities are obtained in our case by generating a template for each pixel being decoded, using the template to generate an integer (the context of the pixel), and using that integer as a pointer to a table of 7.15 JBIG2 571 Name Type Size Signed? Description MMR I 1 N MMR or arithcoding used GBW I 32 N Width of region GBH I 32 N Height of region GBTEMPLATE I 2 N Template number TPON I 1 N Typical prediction used? USESKIP I 1 N Skip some pixels? SKIP B Bitmap for skipping GBATX1 I 8 Y Relative X coordinate of A1 GBATY1 I 8 Y Relative Y coordinate of A1 GBATX2 I 8 Y Relative X coordinate of A2 GBATY2 I 8 Y Relative Y coordinate of A2 GBATX3 I 8 Y Relative X coordinate of A3 GBATY3 I 8 Y Relative Y coordinate of A3 GBATX4 I 8 Y Relative X coordinate of A4 GBATY4 I 8 Y Relative Y coordinate of A4 Table 7.100: Parameters for Decoding a Generic Region. probabilities. Parameter GBTEMPLATE specifies which of four types of templates should be used. Figure 7.101a–d shows the four templates that correspond to GBTEMPLATE values of 0–3, respectively (notice that the two 10-bit templates are identical to templates used by JBIG1, Figure 7.88b,d). The pixel labeled “O” is the one being decoded, and the “X” and the “Ai” are known pixels. If “O” is near an edge of the region, its missing neighbors are assumed to be zero. A4 X X X A3 A2 X X X X X A1 X X X X O X X X X X X X X X A1 X X X O (a) (b) X X X X X X X A1 X X O X X X X X A1 X X X X O (c) (d) Figure 7.101: Four Templates for Generic Region Decoding. The values of the “X” and “Ai” pixels (one bit per pixel) are combined to form a context (an integer) of between 10 and 16 bits. It is expected that the bits will be collected top to bottom and left to right, but the standard does not specify that. The two templates of Figure 7.102a,b, e.g., should produce the contexts 1100011100010111 and 572 7. Image Compression 1000011011001, respectively. Once a context has been computed, it is used as a pointer to a probability table, and the probability found is sent to the arithmetic decoder to decode pixel “O.” 1 1 0 0 0 1 1 1 0 0 0 1 0 1 1 1 O 1 0 0 0 0 1 1 0 1 1 0 0 1 O (a) (b) Figure 7.102: Two Templates. The encoder specifies which of the four template types should be used. A simple rule of thumb is to use large templates for large regions. An interesting feature of the templates is the use of the Ai pixels. They are called adaptive or AT pixels and can be located in positions other than the ones shown. Figure 7.101 shows their normal positions, but a sophisticated encoder may discover, for example, that the pixel three rows above the current pixel “O” is always identical to “O.” The encoder may in such a case tell the decoder (by means of parameters GBATXi and GBATYi) to look for A1 at address (0,-3) relative to “O.” Figure 7.103 shows the permissible positions of the AT pixels relative to the current pixel, and their coordinates. Notice that an AT pixel may also be located at one of the “X” positions of the template. (-128,-128)· · · (-1,-128) (0,-128) (1,-128) · · · (127,-128) ... ... ... (-128,-1) · · · (-1,-1) (0,-1) (1,-1) · · · (127,-1) (-128,-128)· · · (-128,0) (-1,0) O Figure 7.103: Coordinates and Allowed Positions of AT Pixels.  Exercise 7.22: What are the relative coordinates of the AT pixels in the four types of templates shown in Figure 7.101? The TPON parameter controls the so-called typical prediction feature. A typical row is defined as a row that is identical to its predecessor. If the encoder notices that certain rows of the generic region being encoded are typical, it sets TPON to 1 and then writes a code in the compressed file before each row, indicating whether or not it 7.15 JBIG2 573 is typical. If the decoder finds TPON set to 1, it has to decode and check a special code preceding each row of pixels. When the code of a typical row is found, the decoder simply generates it as a copy of its predecessor. The two parameters USESKIP and SKIP control the skip feature. If the encoder discovers that the generic region being encoded is sparse (i.e., most of its pixels are zero), it sets USESKIP to 1 and sets SKIP to a bitmap of size GBWΧGBH, where each 1 bit indicates a zero-bit in the generic region. The encoder then compresses only the 1 bits of the generic region. 7.15.2 Symbol Region Decoding This procedure is invoked when the decoder starts reading a new segment from the compressed file and identifies it as a symbol-region segment. The procedure starts by reading many parameters from the segment. It then inputs the coded information for each symbol and decodes it (using either arithmetic coding or MMR). This information contains the coordinates of the symbol relative to its row and the symbol preceding it, a pointer to the symbol in the symbol dictionary, and possibly also refinement information. The coordinates of a symbol are denoted by S and T. Normally, S is the x-coordinate and T is the y-coordinate of a symbol. However, if parameter TRANSPOSED has the value 1, the meaning of S and T is reversed. In general, T is considered the height of a row of text and S is the coordinate of a text symbol in its row.  Exercise 7.23: What is the advantage of transposing S and T? Symbols are encoded and written on the compressed file by strips. A strip is normally a row of symbols but can also be a column. The encoder makes that decision and sets parameter TRANSPOSED to 0 or 1 accordingly. The decoder therefore starts by decoding the number of strips, then the strips themselves. For each strip the compressed file contains the strip’s T coordinate relative to the preceding strip, followed by coded information for the symbols constituting the strip. For each symbol this information consists of the symbol’s S coordinate (the gap between it and the preceding symbol), its T coordinate (relative to the T coordinate of the strip), its ID (a pointer to the symbol dictionary), and refinement information (optionally). In the special case where all the symbols are aligned vertically on the strip, their T coordinates will be zero. Once the absolute coordinates (x, y) of the symbol in the symbol region have been computed from the relative coordinates, the symbol’s bitmap is retrieved from the symbol dictionary and is combined with the page buffer. However, parameter REFCORNER indicates which of the four corners of the bitmap should be placed at position (x, y). Figure 7.104 shows examples of symbol bitmaps aligned in different ways. If the encoder decides to use MMR to encode a region, it selects one of 15 Huffman code tables defined by the JBIG2 standard and sets a parameter to indicate to the decoder which table is used. The tables themselves are built into both encoder and decoder. Table 7.105 shows two of the 15 tables. The OOB value (out of bound) is used to terminate a list in cases where the length of the list is not known in advance. 7.15.3 Halftone Region Decoding This procedure is invoked when the decoder starts reading a new segment from the compressed file and identifies it as a halftone-region segment. The procedure starts by 574 7. Image Compression • • • Figure 7.104: Three Symbol Bitmaps Aligned at Different Corners. Value Code 0–15 0+Value encoded as 4 bits 16–271 10+(Value - 16) encoded as 8 bits 272–65807 110+(Value - 272) encoded as 16 bits 65808–. 111+(Value - 65808) encoded as 32 bits Value Code 0 0 1 10 2 110 3–10 1110+(Value - 3) encoded as 3 bits 11–74 11110+(Value - 11) encoded as 6 bits 75–. 111110+(Value - 75) encoded as 32 bits OOB 11111 Table 7.105: Two Huffman Code Tables for JBIG2 Decoding. reading several parameters from the segment and setting all the pixels in the halftone region to the value of background parameter HDEFPIXEL. It then inputs the coded information for each halftone pattern and decodes it using either arithmetic coding or MMR. This information consists of a pointer to a halftone pattern in the halftone dictionary. The pattern is retrieved and is logically combined with the background pixels that are already in the halftone region. Parameter HCOMBOP specifies the logical combination. It can have one of the values REPLACE, OR, AND, XOR, and XNOR. Notice that the region information read by the decoder from the compressed file does not include the positions of the halftone patterns. The decoder adds the patterns to the region at consecutive grid points of the halftone grid, which is itself defined by 7.15 JBIG2 575 four parameters. Parameters HGX and HGY specify the origin of the grid relative to the origin of the halftone region. Parameters HRX and HRY specify the orientation of the grid by means of an angle .. The last two parameters can be interpreted as the horizontal and vertical components of a vector v. In this interpretation . is the angle between v and the x axis. The parameters can also be interpreted as the cosine and sine of ., respectively. (Strictly speaking, they are multiples of the sine and cosine, since sin2 . + cos2 . equals unity, but the sum HRX2 + HRY2 can have any value.) Notice that these parameters can also be negative and are not limited to integer values. They are written on the compressed file as integers whose values are 256 times their real values. Thus, for example, if HRX should be -0.15, it is written as the integer -38 because -0.15 Χ 256 = -38.4. Figure 7.106a shows typical relations between a page, a halftone region, and a halftone grid. Also shown are the coordinates of three points on the grid. The decoder performs a double loop in which it varies mg from zero to HGH- 1 and for each value of mg it varies ng from zero to HGW - 1 (parameters HGH and HGW are the number of horizontal and vertical grid points, respectively). At each iteration the pair (ng,mg) constitutes the coordinates of a grid point. This point is mapped to a point (x, y) in the halftone region by the relation x = HGX+mg ΧHRY + ng ΧHRX, y = HGY +mg ΧHRX- ng ΧHRY. (7.27) To understand this relation the reader should compare it to a rotation of a point (ng,mg) through an angle . about the origin. Such a rotation is expressed by (x, y) = (ng,mg)cos . -sin . sin . cos . = (ng cos . +mg sin .,mg cos . - ng sin .). (7.28) A comparison of Equations (7.27) and (7.28) shows that they are identical if we associate HRX with cos . and HRY with sin . (Equation (7.27) also adds the origin of the grid relative to the region, so it results in a point (x, y) whose coordinates are relative to the origin of the region). It is expected that the grid would normally be identical to the region, i.e., its origin would be (0, 0) and its rotation angle would be zero. This is achieved by setting HGX, HGY, and HRY to zero and setting HRX to the width of a halftone pattern. However, all four parameters may be used, and a sophisticated JBIG2 encoder may improve the overall compression quality by trying different combinations. Once point (x, y) has been calculated, the halftone pattern is combined with the halftone region such that its top-left corner is at location (x, y) of the region. Since parts of the grid may lie outside the region, any parts of the pattern that lie outside the region are ignored (see gray areas in Figure 7.106b). Notice that the patterns themselves are oriented with the region, not with the grid. Each pattern is a rectangle with horizontal and vertical sides. The orientation of the grid affects only the position of the top-left corner of each pattern. As a result, the patterns added to the halftone region generally overlap by an amount that’s greatly affected by the orientation and position of the grid. This is also illustrated by the three examples of Figure 7.106b. 576 7. Image Compression (HGX,HGY) Halftone Region One page of the document Halftone Grid (0,0) (HGX+HRX,HGY-HRY) (HGX+HRY,HGY+HRX) x y mg ng (a) (b) Figure 7.106: Halftone Grids and Regions. 7.16 Simple Images: EIDAC 577 7.15.4 The Overall Decoding Process The decoder starts by reading general information for page 1 of the document from the compressed file. This information tells the decoder to what background value (0 or 1) to set the page buffer initially, and which of the four combination operators OR, AND, XOR, and XNOR to use when combining pixels with the page buffer. The decoder then reads segments from the compressed file until it encounters the page information for page 2 or the end of the file. A segment may specify its own combination operator, which overrides the one for the entire page. A segment can be a dictionary segment or an image segment. The latter has one of four types: 1. An immediate direct image segment. The decoder uses this type to decode a region directly into the page buffer. 2. An intermediate direct image segment. The decoder uses this type to decode a region into an auxiliary buffer. 3. An immediate refinement image segment. The decoder uses this type to decode an image and combine it with an existing region in order to refine that region. The region being refined is located in the page buffer, and the refinement process may use an auxiliary buffer (which is then deleted). 4. An intermediate refinement image segment. The decoder uses this type to decode an image and combine it with an existing region in order to refine that region. The region being refined is located in an auxiliary buffer. We therefore conclude that an immediate segment is decoded into the page buffer and an intermediate segment involves an auxiliary buffer. 7.16 Simple Images: EIDAC Image compression methods based on transforms perform best on continuous-tone images. There is, however, an important class of images where transform-based methods, such as JPEG (Section 7.10) and the various wavelet methods (Chapter 8), produce mediocre compression. This is the class of simple images. A simple image is one that uses a small fraction of all the possible grayscales or colors available to it. A common example is a bi-level image where each pixel is represented by eight bits. Such an image uses just two colors out of a palette of 256 possible colors. Another example is a grayscale image scanned from a bi-level image. Most pixels will be black or white, but some pixels may have other shades of gray. A cartoon is also an example of a simple image (especially a cheap cartoon, where just a few colors are used). A typical cartoon consists of uniform areas, so it may use a small number of colors out of a potentially large palette. EIDAC is an acronym that stands for embedded image-domain adaptive compression [Yoo et al. 98]. This method is especially designed for the compression of simple images and combines high compression factors with lossless performance (although lossy compression is an option). The method is also progressive. It compresses each bitplane separately, going from the most-significant bitplane (MSBP) to the least-significant one (LSBP). The decoder reads the data for the MSBP first, and immediately generates a rough “black-and-white” version of the image. This version is improved each time a 578 7. Image Compression bitplane is read and decoded. Lossy compression is achieved if some of the LSBPs are not processed by the encoder. The encoder scans each bitplane in raster order and uses several near neighbors of the current pixel X as the context of X, to determine a probability for X. Pixel X and its probability are then sent to an adaptive arithmetic encoder that does the actual encoding. Thus, EIDAC resembles JBIG and CALIC, but there is an important difference. EIDAC uses a two-part context for each pixel. The first part is an intra context, consisting of several neighbors of the pixel in the same bitplane. The second part is an inter context, whose pixels are selected from the already encoded bitplanes (recall that encoding is done from the MSBP to the LSBP). To see why an inter context makes sense, consider a grayscale image scanned from a bi-level image. Most pixel values will be 255 (1111 11112) or 0 (0000 00002), but some pixels may have values close to these, such as 254 (1111 11102) or 1 (0000 00012). Figure 7.107a shows (in binary) the eight bitplanes of the six pixels 255 255 254 0 1 0. It is clear that there are spatial correlations between the bitplanes. A bit position that has a zero in one bitplane tends to have zeros in the other bitplanes. Another example is a pixel P with value “0101xxxx” in a grayscale image. In principle, the remaining four bits can have any of 16 different values. In a simple image, however, the remaining four bits will have just a few values. There is therefore a good chance that pixels adjacent to P will have values similar to P and will therefore be good predictors of P in all the bitplanes. BP7 1 1 1 0 0 0 . . . BP6 1 1 1 0 0 0 . . . BP5 1 1 1 0 0 0 . . . ... ... ... ... ... ... . . . BP0 1 1 0 0 1 0 . . . a b c d X MSBP current BP LSBP a b c d d a b c X (a) (b) (c) Figure 7.107: Pixel Contexts Used by EIDAC. The intra context actually used by EIDAC is shown in Figure 7.107b. It consists of four of the already-seen nearest neighbors of the current pixel X. The inter context is shown in Figure 7.107c. It consists of (1) the same four neighbors in the bitplane immediately above the current bitplane, (2) the same four neighbors in the MSBP, and (3) the pixels in the same position as X in all the bitplanes above it (five shaded pixels in the figure). Thus, a pixel in the MSBP has no inter context, while the inter context of a pixel in the LSBP consists of 4+4+7 = 15 pixels. Other contexts can be selected, which means that a general implementation of EIDAC should include three items of side information in the compressed stream. The first item is the dimensions of the image, the second item is the way the contexts are selected, and the third item is a flag indicating whether histogram compaction is used (if it is used, the flag should be followed by the new codes). This useful feature is described here. 7.17 Block Matching 579 A simple image is one that has a small number of colors or grayscales out of a large available palette of colors. Imagine a simple image with eight bits per pixel. If the image has just 27 colors, each can be assigned a 5-bit code instead of the original 8-bit value. This feature is referred to as histogram compaction. When histogram compaction is used, the new codes for the colors have to be included in the compressed stream, where they constitute overhead. Generally, histogram compaction improves compression, but in rare cases the overhead may be bigger than the savings due to histogram compaction. 7.17 Block Matching The LZ77 sliding window method for compressing text (Section 6.3) can be applied to images as well. This section describes such an application for the lossless compression of images. It follows the ideas outlined in [Storer and Helfgott 97]. Figure 7.108a shows how a search buffer and a look-ahead buffer can be made to slide in raster order along an image. The principle of image compression (Section 7.3) says that the neighbors of a pixel P tend to have the same value as P or very similar values. Thus, if P is the leftmost pixel in the look-ahead buffer, there is a good chance that some of its neighbors on the left (i.e., in the search buffer) will have the same value as P. There is also a chance that a string of neighbor pixels in the search buffer will match P and the string of pixels that follow it in the look-ahead buffer. Experience also suggests that in any nonrandom image there is a good chance of having identical strings of pixels in several different locations in the image. Since near-neighbors of a pixel are located also above and below it and not just on the left and right, it makes sense to use “wide” search and look-ahead buffers and to compare blocks of, say, 4Χ4 pixels instead of individual pixels (Figure 7.108b). This is the reason for the name block matching. When this method is used, the number of rows and columns of the image should be divisible by 4. If the image doesn’t satisfy this, up to three artificial rows and/or columns should be added. Two reasonable edge conventions are (1) add columns of zero pixels on the left and rows of zero pixels on the top of the image, (2) duplicate the rightmost column and the bottom row as many times as needed. Figure 7.108c,d,e shows three other ways of sliding the window in the image. The -45.-diagonal traversal visits pixels in the order (1,1), (1,2), (2,1), (1,3), (2,2), (3,1),. . . . In the rectilinear traversal pixels are visited in the order (1,1), (1,2), (2,2), (2,1), (1,3), (2,3), (3,3), (3,2), (3,1),. . . . The circular traversal visits pixels in order of increasing distance from the center. The encoder proceeds as in the LZ77 algorithm. It finds all blocks in the search buffer that match the current 4Χ4 block in the look-ahead buffer and selects the longest match. A token is emitted (i.e., written on the compressed stream), and both buffers are advanced. In the original LZ77 the token consists of an offset (distance), match length, and next symbol in the look-ahead buffer (and the buffers are advanced by one plus the length of the match). The third item guarantees that progress will be made even in cases where no match was found. When LZ77 is extended to images, the third item of a token is the next pixel block in the look-ahead buffer. The method should therefore 580 7. Image Compression (a) (b) (c) (d) (e) Search buffer Search buffer Search buffer Look-ahead Look-ahead buffer Look-ahead buffer Figure 7.108: Block Matching (LZ77 for Images). be modified to output tokens without this third item, and three ways of doing this are outlined here.  Exercise 7.24: What’s wrong with including the next block of pixels in the token? 1. The output stream consists of 2-field tokens (offset, length). When no match is found, a token of the form (0,0) is output, followed by the raw unmatched block of pixels (this is common at the start of the compression process, when the search buffer is empty or almost so). The decoder can easily mimic this. 2. The output stream consists of variable-size tags, some of which are followed by either a 2-field token or by a raw block of 16 pixels. This method is especially appropriate for bi-level images, where a pixel is represented by a single bit. We know from experience that in a “typical” bi-level image there are more uniform white areas than uniform black ones (the doubting reader should check the eighth of the ITU-T fax training documents, listed in Table 5.30, for an example of an atypical image). The tag can therefore be assigned one of the following four values: 0 if the current block in the look-ahead buffer 7.17 Block Matching 581 is all white, 10 if the current block is all black, 110 if a match was found (in which case the tag is followed by an offset and a length, encoded in either fixed size or variable size), and 111 if no match was found (in this case the tag is followed by the 16-bit raw pixel block). 3. This is similar to 2 above, but on discovering a current all-white or all-black pixel block, the encoder looks for a run of such blocks. If it finds n consecutive all-white blocks, it generates a tag of 0 followed by the encoded value of n. Similarly, a tag of 10 precedes an encoded count of a run of all-black pixel blocks. Since large values of n are rare, it makes sense to use a variable-length code such as the general unary codes of Section 3.1. A simple encoder can always use the same unary code, for example, the (2,1,10) code, which has 2044 values. A sophisticated encoder may use the image resolution to estimate the maximum size of a run of identical blocks, select a proper value for k based on this, and use code (2, 1, k). The value of k being used would have to be included in the header of the compressed file, for the decoder’s use. Table 7.109 lists the number of codes of each of the general unary codes (2, 1, k) for k = 2, 3, . . . , 11. This table was calculated by the single Mathematica statement Table[2^(k+1)-4,{k,2,11}]. k : 2 3 4 5 6 7 8 9 10 11 (2, 1, k): 4 12 28 60 124 252 508 1020 2044 4092 Table 7.109: Number of General Unary Codes (2, 1, k) for k = 2, 3, . . . , 11. The offset can be a single number, the distance of the matched string of pixel blocks from the edge of the search buffer. This has the advantage that the maximum value of the distance depends on the size of the search buffer. However, since the search buffer may meander through the image in a complex way, the decoder may have to work hard to determine the image coordinates from the distance. An alternative is to use as offset the image coordinates (row and column) of the matched string instead of its distance. This makes it easy for the decoder to find the matched string of pixel blocks, but it decreases compression performance, since the row and column may be large numbers and since the offset does not depend on the actual distance (offsets for a short distance and for a long distance would require the same number of bits). An example may shed light on this point. Imagine an image with a resolution of 1KΧ1K. The numbers of rows and columns are 10 bits each, so coding an offset as the pair (row, column) requires 20 bits. A sophisticated encoder, however, may decide, based on the image size, to select a search buffer of 256K = 218 locations (25% of the image size). Coding the offsets as distances in this search buffer would require 18 bits, a small gain. 7.17.1 Implementation Details The method described here for finding a match is just one of many possible methods. It is simple and fast, but is not guaranteed to find the best match. It is similar to the method used by LZP (Section 6.19) to find matches. Notice that finding matches is a task performed by the encoder. The decoder’s task is to use the token to locate a string of pixel blocks, and this is straightforward. 582 7. Image Compression The encoder starts with the current pixel block B in the look-ahead buffer. It hashes the values of the pixels of B to create a pointer p to an index table T. In T[p] the encoder normally finds a pointer q pointing to the search buffer S. Block B is compared with S[q]. If they are identical, the encoder finds the longest match. If B and S[q] are different, or if T[p] is invalid (does not contain a pointer to the search buffer), there is no match. In any of these cases, a pointer to B is stored in T[p], replacing its former content. Once the encoder finds a match or decides that there is no match, it encodes its findings using one of the methods outlined above. The index table T is initialized to all invalid entries, so initially there are no matches. As more and more pointers to pixel blocks B are stored in T, the index table becomes more useful, since more of its entries contain meaningful pointers. Another implementation detail has to do with the header of the compressed file. It should include side information that depends on the image or on the particular method used by the encoder. Such information is obviously needed by the decoder. The main items included in the header are the image resolution, number of bitplanes, the block size (4 in the examples above), and the method used for sliding the buffers. Other items that may be included are the sizes of the search and look-ahead buffers (they determine the sizes of the offset and the length fields of a token), the edge convention used, the format of the compressed file (one of the three methods outlined above), and whether the offset is a distance or a pair of image coordinates,  Exercise 7.25: Indicate one more item that a sophisticated encoder may have to include in the header. 7.18 Grayscale LZ Image Compression The dictionary-based compression methods described in Chapter 6 are one-dimensional. They were developed mostly for the compression of text, which is one-dimensional (although they do not perform badly on images). Images, however, are two-dimensional, and the algorithm described here is an extension of the traditional LZ methods to two dimensions. This method has been created specifically for the lossless compression of grayscale images and it is an LZ method, i.e., it is based on a dictionary of alreadyencoded data. This is why the name of this method is GS-2D-LZ. The main references are [Brittain and El-Sakka 05 and 07]. The algorithm scans the input image in raster order, encoding blocks of pixels as it goes along. In each step it tries to match a block E of yet-unencoded pixels to a block M of already-encoded pixels. The main novelty of the algorithm is that the matches do not have to be exact. Corresponding pixels in the two matched blocks are generally different, but the differences are within a threshold parameter. The differences (referred to as residuals) are small numbers and are efficiently encoded with an entropy encoder (the actual implementation employs PAQ6, Section 5.15, for this purpose). Figure 7.110 illustrates the principle. The dark squares indicate encoded pixels. The Χ marks the current encoder’s position. The white squares indicate unencoded pixels and the light-gray squares constitute the search region. This region is a rectangle of search-heightΧsearch-width pixels whose bottom-right corner is located at the encoder’s 7.18 Grayscale LZ Image Compression 583 position. Both search-height and search-width are small parameters whose best values need to be determined experimentally. Since they are small, it is reasonable to expect that the pixels inside the search region will be correlated with pixel Χ and its neighbors to the right and below.                       Figure 7.110: Matching a Region. Figure 7.110 assumes that the rectangular 4 Χ 3 block M with black triangles has been (approximately) matched with the same-size block E whose top-left corner is at Χ. The latter block is compressed by encoding (1) the location of the match relative to the block being encoded (five steps to the left and two steps up), (2) the dimensions of the match (4Χ3), and (3) the differences (residuals) between each pixel in the match block M and the corresponding pixel in the encoded block E. Once this is done, the encoder marks the 12 pixels of E as encoded and skips to the pixel marked +. The case where no match can be found in the search region is covered below. Now for the details of (approximate) matching. The first encoder step is to scan the S def = (search-heightΧsearch-width)-1 pixels in the search region and compare each to Χ. If the difference between Χ and a pixel P is less than or equal to the threshold parameter, the encoder considers P an anchor and tries to match a maximal-size block M whose top-left corner is at P to the same-size block E whose top-left corner is Χ. Once all S pixels in the search block have been checked in this way, the largest match M is selected and is used to encode block E as shown above. For each anchor pixel P in the search region, the encoder compares P to Χ. If their difference is within parameter threshold, the encoder compares the pixel to the right of P with the pixel to the right of Χ. This is repeated until a pixel pair is found whose difference is greater than threshold. The encoder has now matched a sequence (a short row) of pixels. It moves down one row and checks the pixel pairs below those that have been matched. The next row may feature the same number of matches or fewer, but the encoder does not try to match a longer sequence. The process repeats row by row until a row is found where no pairs match. At this point, the encoder selects the largest rectangle in the matched region and considers it the matching block M found for anchor pixel P. After all S anchors have been examined in this way, the largest match block M is selected by the encoder and is employed to encode the corresponding block E. This 584 7. Image Compression          !       "      "  !    !   "   Figure 7.111: Matching a Region For An Anchor Pixel P. process is illustrated in Figure 7.111a where the pixel pairs are numbered according to the search order. The figure assumes that anchor pixel P has matched pixel Χ. The encoder managed to match a sequence of seven pixels to the right of P (eight pixels including P), but only seven on the next two rows. Then came two rows of 3-pixel sequences and two more rows of short, 2-pixel sequences. The bold slashes indicate no match. No pixel matched on the next row, thereby signaling the end of the pixel-match loop. Two rectangles can be identified in the matched region, a horizontal 3 Χ 7 rectangle (shown in dashed) and a vertical 7 Χ 2 rectangle. The former is larger, so it is selected (if it satisfies the two rules below) as the match block M for anchor pixel P. Once a match block M has been identified for each of the 6 Χ 6 - 1 = 35 anchor pixels in the search region, the largest M that satisfies the two rules below is selected. Assuming that this is the 3Χ7 rectangle shown in Figure 7.111a in dashed, it is used to encode the 3 Χ 7 rectangle E whose top-left corner is at Χ (shown in bold dashed-dot). Once the differences between the 21 pixel pairs have been encoded, the encoder skips to the pixel indicated by +. Matching involves two more rules. Rule 1. A block M can be a match to a block E only if the number of new pixels in E exceeds the value of parameter minimum-matchsize. Figure 7.111b shows an example where block E contains four already-encoded pixels (marked DC, or don’t care). Those pixels are not counted as new and also do not have to match their opposite numbers in M. Rule 2. The mean square error (MSE, Section 7.4.2) of blocks M and E must be less than the value of parameter max-MSE. It is important to consider the case where no match is found. This may happen if the image region above and to the left of the encoder’s position Χ is very different from the region below Χ and to its right. If the encoder cannot find a match block M for any anchor pixel, it selects a block E of dimensions (no-match-block-height Χ no-match-block-width) and with its top-left corner at Χ. This block is then encoded with a simple prediction method, similar to the one used by CALIC (Section 7.28). 7.18 Grayscale LZ Image Compression 585 Both the context and the prediction rules are shown in Figure 7.112. This ensures that progress is made even in noisy image regions where pixels are not correlated. ## # $# $ $$ $ $$ Dh = |W-WW| + |N -NW| + |NE - N| Dv = |W-NW| + |N - NN| + |NE - NNE| DiffDhDv = Dh - Dv If (DiffDhDv > 80) Return N If (DiffDhDv < -80) Return W Pred = (N + W)/2 + (NE - NW)/4 If (DiffDhDv > 32) Return(Pred + N)/2 If (DiffDhDv < -32) Return(Pred + W)/2 If (DiffDhDv > 8) Return(3 Χ Pred + N)/4 If (DiffDhDv < -8) Return(3 Χ Pred + W)/4 Return(Pred) Figure 7.112: Context And Rules For Predicting Pixels. The main encoder loop of matching pixel pairs and selecting the best matching blocksM can lead to another complication, which is illustrated in Figure 7.113. We know that a data compressor can use only quantities that will be known to its decompressor at the time of decoding. The figure, however, shows a match that has extended to unencoded pixels. These pixels will be unknown to the decoder when it tries to decode pixel Χ, but the surprising fact is that the decoder can handle this case recursively. Figure 7.113 shows a 3 Χ 4 matching block M (in dashed) four of whose pixels are also part of the matched block E (in bold dashed-dot). When the decoder first tries to decode E, it inputs the encoded differences between the 12 pixel pairs of M and E. The decoder first decodes the differences and then adds each difference to a pixel in M to obtain the corresponding pixel in E. In this way, it uses pixels 1–4 to decode pixels 7–10, pixels 5–8 to decode pixels 13–16, and pixels 11–14 to decode 17–20. The decoder now has the values of pixels 7, 8, 13, and 14, and it uses them to decode pixels 15, 16, 19, and 20. This neat process is the two-dimensional extension of LZW decoding (see exercise 6.10). Notice that this type of recursive decoding is possible only if the “problematic” pixels are located to the right and/or below the encoder’s position.     "  !    "  !      Figure 7.113: Decoding With Yet-Unencoded Pixels. 586 7. Image Compression Once the encoder has decided on a match M to a block E (or discovered that there is no match at the current position), it prepares the data needed by the decoder and stores it in five tables as follows: 1. The match-flag table contains a boolean value for each encoding step, true, if a matching block was found; false, in cases of no match. 2. The match-location table, where the position of the matching block M relative to the encoded block E, is recorded. Thus, for the match shown in Figure 7.111b, the match location has coordinates (2, 5), while for the match shown in Figure 7.110, the match location has coordinates (5, 2). 3. The match-dimensions table contains the dimensions of the two matched blocks. 4. The differences between the pixel pairs of blocks M and E are recorded in table residuals. 5. An additional table, prediction errors, is used for those cases where no match was found (i.e., the entry in match-flag is false). Here are stored the differences between predicted and actual pixel values for each pixel in the block. Data is prepared by the encoder and is appended to each table in each encoding step. Once the entire image has been scanned and encoded, each table is in turn encoded with an entropy encoder. After trying various statistical encoders, the developers of GS- 2D-LZ decided to employ PAQ6 (Section 5.15) for this task. Parameter settings The algorithm depends on seven parameters whose values had to be determined experimentally for various types of data. It is intuitively clear that wrong values, whether too high or too low, can adversely affect the performance of the algorithm, and it is enlightening to see how. 1. The dimensions of a match region M are search-width and search-height. Large regions imply better approximate match, because the large number of matching pixel pairs normally results in lower MSE. However, it takes more bits to encode the dimensions of large regions. 2. The threshold parameter restricts the maximum allowable difference between the two pixels of a pair. Large values of this parameter may result in bigger match blocks, at the price of larger residuals and poorer compression. 3. The maximum MSE between blocks is specified by the max-MSE parameter. Small values of the parameter result in closer matches, but also in smaller matched blocks and more cases of no match. 4. The minimum number of new pixels in a matched block E is specified by parameter min-match-size. Large values imply larger matched blocks, but also fewer matches and more cases of no match. 5. Finally, parameters no-match-block-width and no-match-block-height determine the dimensions of blocks in cases of no match. Large values cause faster progress in such cases, but a smaller chance of employing the LZ principle for finding matches. Experiments to determine the optimal values of the parameters were run with 24 images as test data. The images were divided into four classes—natural scenes (images of people, houses, birds, and a hill), geographic (taken from high altitudes, normally 7.18 Grayscale LZ Image Compression 587 a satellite), graphic (hand drawn or computer generated), and medical (X-rays, ultrasound, and magnetic resonance)—with six images in each class. The results are summarized in Table 7.114 and show a clear preference of small blocks (typically 4 Χ 4 or 5 Χ 5) and large thresholds (about 25 out of 256 possible shades of grey). Natural scene Geographic Graphic Medical search-width 4 5 5 4 search-height 4 5 5 4 threshold 27 25 25 27 min-match-size 17 17 21 16 max-MSE 2.5 3.0 3.8 1.8 no-match-block-width 5 5 5 5 no-match-block-height 5 5 5 5 Table 7.114: Optimal Settings For Seven Parameters. Finally, with the parameters set to their optimal values, the algorithm was tested on 112 grayscale images. It was found to perform as well or even outperform popular LZ methods (Table 7.115) and even state-of-the-art algorithms such as JPEG-LS, the lossless option of JPEG2000, and CALIC (Table 7.116). Image class GS-2D-LZ PNG GIF UNIX Compress Natural scene 4.50 4.73 6.83 6.21 Geographic 5.17 5.40 7.37 6.46 Graphic 1.61 1.77 2.73 2.48 Medical 3.25 3.42 4.96 4.58 Table 7.115: GS-2D-LZ Compared to Popular LZ Codecs. Image name GS-2D-LZ BZIP2 JPEG2000 JPEG-LS CALIC-h CALIC-a Natural scene 4.50 4.88 4.66 4.71 4.50 4.27 Geographic 5.17 5.24 5.31 5.24 5.23 5.06 Graphic 1.61 1.77 2.61 1.90 2.11 1.73 Medical 3.25 3.48 3.22 3.22 3.39 3.23 Table 7.116: GS-2D-LZ Compared to State-Of-The-Art Codecs. I would like to thank Prof. El-Sakka for his help. He has read this section and provided useful comments and corrections. 588 7. Image Compression 7.19 Vector Quantization Vector quantization is a generalization of the scalar quantization method (Section 1.6). It is used for both image and audio compression. In practice, vector quantization is commonly used to compress data that has been digitized from an analog source, such as audio samples and scanned images. Such data is called digitally sampled analog data (DSAD). Vector quantization is based on the fact that compression methods that compress strings (or blocks) rather than individual symbols (Section 6.1) can, in principle, produce better results. Such methods are called block coding and, as the block length increases, compression methods based on block coding can approach the entropy when compressing stationary information sources. We start with a simple, intuitive vector quantization method for image compression. Given an image, we divide it into small blocks of pixels, typically 2 Χ 2 or 4Χ4. Each block is considered a vector. The encoder maintains a list (called a codebook) of vectors and compresses each block by writing on the compressed stream a pointer to the block in the codebook. The decoder has the easy task of reading pointers, following each pointer to a block in the codebook, and appending the block to the image-so-far. Thus, vector quantization is an asymmetric compression method. In the case of 2Χ2 blocks, each block (vector) consists of four pixels. If each pixel is one bit, then a block is four bits long and there are only 24 = 16 different blocks. It is easy to store such a small, permanent codebook in both encoder and decoder. However, a pointer to a block in such a codebook is, of course, four bits long, so there is no compression gain by replacing blocks with pointers. If each pixel is k bits, then each block is 4k bits long and there are 24k different blocks. The codebook grows very fast with k (for k = 8, for example, it is 2564 = 232 = 4 Tera entries) but the point is that we again replace a block of 4k bits with a 4k-bit pointer, resulting in no compression gain. This is true for blocks of any size. Once it becomes clear that this simple method does not work, the next thing that comes to mind is that any given image may not contain every possible block. Given 8-bit pixels, the number of 2Χ2 blocks is 22·2·8 = 232 . 4.3 billion, but any particular image may contain only a few million pixels and a few thousand different blocks. Thus, our next version of vector quantization starts with an empty codebook and scans the image block by block. The codebook is searched for each block. If the block is already in the codebook, the encoder outputs a pointer to the block in the (growing) codebook. If the block is not in the codebook, it is added to the codebook and a pointer is output. The problem with this simple method is that each block added to the codebook has to be written on the compressed stream. This greatly reduces the effectiveness of the method and may lead to low compression and even to expansion. There is also the small added complication that the codebook grows during compression, so the pointers get longer, but this is not difficult for the decoder to handle. These problems are the reason why image vector quantization is lossy. If we accept lossy compression, then the size of the codebook can be greatly reduced. Here is an intuitive lossy method for image compression by vector quantization. Analyze a large number of different “training” images and find the B most-common blocks. Construct a codebook with these B blocks and embed it into both encoder and decoder. Each entry of the codebook is a block. To compress an image, scan it block by block, and for each 7.19 Vector Quantization 589 block find the codebook entry that best matches it, and output a pointer to that entry. The size of the pointer is, of course, log2 B, so the compression ratio (which is known in advance) is log2 B block size . One problem with this approach is how to match image blocks to codebook entries. Here are a few common measures. Let B = (b1, b2, . . . , bn) and C = (c1, c2, . . . , cn) be a block of image pixels and a codebook entry, respectively (each is a vector of n bits). We denote the “distance” between them by d(B,C) and measure it in three different ways as follows: d1(B,C) = n i=0 |bi - ci|, d2(B,C) = n i=0 (bi - ci)2, (7.29) d3(B,C) = MAXn i=0|bi - ci|. The third measure d3(B,C) is easy to interpret. It finds the component where B and C differ most, and it returns this difference. The first two measures are easy to visualize in the case n = 3. Measure d1(B,C) becomes the distance between the two threedimensional vectors B and C when we move along the coordinate axes. Measure d2(B,C) becomes the Euclidean (straight line) distance between the two vectors. The quantities di(B,C) can also be considered measures of distortion. A distortion measure (often also called a metric) is a function m that measures “distance” between, or “closeness” of, two mathematical quantities a and b. Such a function should satisfy m(a, b) . 0, m(a, a) = 0, and m(a, b) +m(b, c) . m(a, c). Another problem with this approach is the quality of the codebook. In the case of 2Χ2 blocks with 8-bit pixels the total number of blocks is 232 . 4.3 billion. If we decide to limit the size of our codebook to, say, a million entries, it will contain only 0.023% of the total number of blocks (and still be 32 million bits, or about 4 Mb long). Using this codebook to compress an “atypical” image may result in a large distortion regardless of the distortion measure used. When the compressed image is decompressed, it may look so different from the original as to render our method useless. A natural way to solve this problem is to modify the original codebook entries in order to adapt them to the particular image being compressed. The final codebook will have to be included in the compressed stream, but since it has been adapted to the image, it may be small enough to yield a good compression ratio, yet close enough to the image blocks to produce an acceptable decompressed image. Such an algorithm has been developed by Linde, Buzo, and Gray [Linde, Buzo, and Gray 80]. It is known as the LBG algorithm and it is the basis of many vector quantization methods for the compression of images and sound (see, for example, Section 7.37). Its main steps are the following: Step 0: Select a threshold value  and set k = 0 and D(-1) = .. Start with an initial codebook with entries C(k) i (where k is currently zero, but will be incremented in each 590 7. Image Compression iteration). Denote the image blocks by Bi (these blocks are also called training vectors, since the algorithm uses them to find the best codebook entries). Step 1: Pick up a codebook entry C(k) i . Find all the image blocks Bm that are closer to Ci than to any other Cj . Phrased more precisely; find the set of all Bm that satisfy d(Bm,Ci) < d(Bm,Cj) for all j = i. This set (or partition) is denoted by P(k) i . Repeat for all values of i. It may happen that some partitions will be empty, and we deal with this problem below. Step 2: Select an i and calculate the distortion D(k) i between codebook entry C(k) i and the set of training vectors (partition) P(k) i found for it in step 1. Repeat for all i, then calculate the average D(k) of all the D(k) i . A distortion D(k) i for a certain i is calculated by computing the distances d(C(k) i ,Bm) for all the blocks Bm in partition P(k) i , then computing the average distance. Alternatively, D(k) i can be set to the minimum of the distances d(C(k) i ,Bm). Step 3: If (D(k-1) - D(k))/D(k) . , halt. The output of the algorithm is the last set of codebook entries C(k) i . This set can now be used to (lossy) compress the image with vector quantization. In the first iteration k is zero, so D(k-1) = D(-1) = . > . This guarantees that the algorithm will not stop at the first iteration. Step 4: Increment k by 1 and calculate new codebook entries C(k) i ; each equals the average of the image blocks (training vectors) in partition P(k-1) i that was computed in step 1. (This is how the codebook entries are adapted to the particular image.) Go to step 1. A full understanding of such an algorithm calls for a detailed example, parts of which should be worked out by the reader in the form of exercises. In order to easily visualize the example, we assume that the image to be compressed consists of 8-bit pixels, and we divide it into small, two-pixel blocks. Normally, a block should be square, but the advantage of our two-pixel blocks is that they can be plotted on paper as twodimensional points, thereby rendering the data (as well as the entire example) more visual. Examples of blocks are (35, 168) and (250, 37); we interpret the two pixels of a block as the (x, y) coordinates of a point. Our example assumes an image consisting of 24 pixels, organized in the 12 blocks B1 = (32, 32), B2 = (60, 32), B3 = (32, 50), B4 = (60, 50), B5 = (60, 150), B6 = (70, 140), B7 = (200, 210), B8 = (200, 32), B9 = (200, 40), B10 = (200, 50), B11 = (215, 50), and B12 = (215, 35) (Figure 7.117). It is clear that the 12 points are concentrated in four regions. We select an initial codebook with the four entries C(0) 1 = (70, 40), C(0) 2 = (60, 120), C(0) 3 = (210, 200), and C(0) 4 = (225, 50) (shown as x in the diagram). These entries were selected more or less at random but we show later how the LBG algorithm selects them methodically, one by one. Because of the graphical nature of the data, it is easy to determine the four initial partitions. They are P(0) 1 = (B1,B2,B3,B4), P(0) 2 = (B5,B6), P(0) 3 = (B7), and P(0) 4 = (B8,B9,B10,B11,B12). Table 7.118 shows how the average distortion D(0) is calculated for the first iteration (we use the Euclidean 7.19 Vector Quantization 591 B1 B2 B3 B4 B5 C2 C1 C3 C4 B6 B7 B8 B9 B10 B11 B12 Χ Χ Χ Χ 20 20 40 40 60 60 80 80 100 100 120 120 140 140 160 160 180 180 200 200 220 220 240 240 Figure 7.117: Twelve Points and Four Codebook Entries C(0) i . I: (70 - 32)2 + (40 - 32)2 = 1508, (70 - 60)2 + (40 - 32)2 = 164, (70 - 32)2 + (40 - 50)2 = 1544, (70 - 60)2 + (40 - 50)2 = 200, II: (60 - 60)2 + (120 - 150)2 = 900, (60 - 70)2 + (120 - 140)2 = 500, III: (210 - 200)2 + (200 - 210)2 = 200, IV: (225 - 200)2 + (50 - 32)2 = 449, (225 - 200)2 + (50 - 40)2 = 725, (225 - 200)2 + (50 - 50)2 = 625, (225 - 215)2 + (50 - 50)2 = 100, (225 - 215)2 + (50 - 35)2 = 325. Table 7.118: Twelve Distortions for k = 0. 592 7. Image Compression distance function). The result is D(0) =(1508 + 164 + 1544 + 200 + 900 + 500 + 200 + 449 + 725 + 625 + 100 + 325)/12 =603.33. Step 3 indicates no convergence, since D(-1) = ., so we increment k to 1 and calculate four new codebook entries C(1) i (rounded to the nearest integer for simplicity) C(1) 1 = (B1 + B2 + B3 + B4)/4 = (46, 41), C(1) 2 = (B5 + B6)/2 = (65, 145), C(1) 3 = B7 = (200, 210), C(1) 4 = (B8 + B9 + B10 + B11 + B12)/5 = (206, 41). (7.30) They are shown in Figure 7.119.  Exercise 7.26: Perform the next iteration.  Exercise 7.27: Use the four codebook entries of Equation (7.30) to perform the next iteration of the LBG algorithm. A slightly different interpretation of the LBG algorithm can provide intuitive understanding of why it works. Imagine a vector quantizer that consists of an encoder and a decoder. For a given input vector, the encoder finds the region with the most similar samples, the decoder outputs for a given region, the most representative vector for that region. The LBG algorithm optimizes the encoder (in step 1) and the decoder (in step 4) until no further improvement is possible. Since the algorithm strictly reduces the average distortion from one iteration to the next, it does not guarantee that the codebook entries will converge to the optimum set. In general, they will converge to a local minimum, a miminum that can be improved only by temporarily making the distortion of the solution worse (which we know is impossible for LBG). The quality of the solution and the speed of convergence depend heavily on the initial choice of codebook entries (i.e., on the values of C(0) i ). We therefore discuss this aspect of the LBG algorithm next. The original LBG algorithm proposes a computationally intensive splitting technique where the initial codebook entries C(0) i are selected in several steps as follows: Step 0: Set k = 1 and select a codebook with k entries (i.e., one entry) C(0) 1 that’s an average of all the image blocks Bm. Step 1: In a general iteration there will be k codebook entries C(0) i , i = 1, 2, . . . , k. Split each entry (which is a vector) into the two similar entries C(0) i ± e where e is a fixed perturbation vector. Set k ‹ 2k. (Alternatively, we can use the two vectors C(0) i and C(0) i + e, a choice that leads to smaller overall distortion.) Step 2: If there are enough codebook entries, stop the splitting process. The current set of k codebook entries can now serve as the initial set C(0) i for the LBG algorithm above. 7.19 Vector Quantization 593 B1 B2 B3 B4 B5 C2 C1 C3 C4 B6 B7 B8 B9 B10 B11 B12 Χ Χ Χ Χ 20 20 40 40 60 60 80 80 100 100 120 120 140 140 160 160 180 180 200 200 220 220 240 240 Figure 7.119: Twelve Points and Four Codebook Entries C(1) i . I: (46 - 32)2 + (41 - 32)2 = 277, (46 - 60)2 + (41 - 32)2 = 277, (46 - 32)2 + (41 - 50)2 = 277, (46 - 60)2 + (41 - 50)2 = 277, II: (65 - 60)2 + (145 - 150)2= 50, (65 - 70)2 + (145 - 140)2= 50, III: (210 - 200)2 + (200 - 210)2 = 200, IV: (206 - 200)2 + (41 - 32)2 = 117, (206 - 200)2 + (41 - 40)2= 37, (206 - 200)2 + (41 - 50)2 = 117, (206 - 215)2 + (41 - 50)2 = 162, (206 - 215)2 + (41 - 35)2 = 117. Table 7.120: Twelve Distortions for k = 1. 594 7. Image Compression If more entries are needed, execute the LBG algorithm on the current set of k entries, to converge them to a better set; then go to step 1. This process starts with one codebook entry C(0) 1 and uses it to create two entries, then four, eight, and so on. The number of entries is always a power of 2. If a different number is needed, then the last time step 1 is executed, it can split just some of the entries. For example, if 11 codebook entries are needed, then the splitting process is repeated until eight entries have been computed. Step 1 is then invoked to split just three of the eight entries, ending up with 11 entries. One more feature of the LBG algorithm needs to be clarified. Step 1 of the algorithm says; “Find all the image blocks Bm that are closer to Ci than to any other Cj . They become partition P(k) i .” It may happen that no image blocks are close to a certain codebook entry Ci, which creates an empty partition P(k) i . This is a problem, because the average of the image blocks included in partition P(k) i is used to compute a better codebook entry in the next iteration. An empty partition is also bad because it “wastes” a codebook entry. The distortion of the current solution can be trivially reduced by replacing codebook entry Ci with any image block. An even better solution is to delete codebook entry Ci and replace it with a new entry chosen at random from one of the image blocks included in the partition P(k) j having biggest distortion. If an image block is an n-dimensional vector, then the process of constructing the partitions in step 1 of the LBG algorithm divides the n-dimensional space into Voronoi regions [Salomon 99] with a codebook entry Ci at the center of each region. Figures 7.118 and 7.119 show the Voronoi regions in the first two iterations of our example. In each iteration the codebook entries are moved, thereby changing the Voronoi regions. Speeding up the codebook search: In order to quantize V = (v1, v2, . . . , vn) with a vector quantizer having a k-entry codebook, the method described above requires the comparison of the vector V with each entry in the codebook. Each comparison involves the computation of the distortion between the vector and a codeword. This approach to vector quantization is called Exhaustive Search Vector Quantizer or ESVQ. An exhaustive search of high-dimensional vectors in a large codebook can be computationally prohibitive, so several algorithms have been designed to overcome this issue. A divide and conquer search algorithm, very similar to the well-known binary search, can speed up the codebook search. For simplicity, we describe this method for the two-dimensional vector quantizer presented in Figure 7.119, however, it is possible to generalize it to quantizers of any dimension. Line A in Figure 7.121 separates the two-dimensional space containing the codewords into two semi-planes A+ and A-. A similar subdivision is performed by lines B through E. For reasons that will become clear later, all lines in the figure mark one or more boundaries of the quantizer regions. If, instead of comparing vector V to each codeword, we check whether V lies to the left or to the right of a line, we can (almost) halve the size of the search space with a single comparison. By repeating the comparison with the other lines and intersecting the resulting semi-planes, we can pinpoint the region to which V belongs, and thus also the closest codeword in the codebook. The sequence of decisions, illustrated in the form of a decision tree, is depicted in Figure 7.122. For example, to quantize V , we start comparing it to line A. V lies on the semi-plane A+, so we proceed by comparing V to line B. Since V lies on the 7.19 Vector Quantization 595                                            Figure 7.121: Quantizer Regions Partitioned into Semi-Planes. semi-plane B+, we continue by comparing V to C and conclude that V lies on the semiplane C+. The intersection of the semi-planes A+, B+, and C+ is the quantizer region corresponding to codeword C4. From the depth of the decision tree it is evident, even with such a small example, that the number of comparisons is reduced from four to at most three. If the vector being quantized lies in the region associated with C3, then only two comparisons are necessary. It is important to notice that this method speeds up the codeword search without degrading the quality of the result. In the following paragraphs we introduce methods that are capable of even faster searches while marginally compromising the overall distortion of the quantizer. All the methods described impose some sort of structure on 596 7. Image Compression                 Figure 7.122: Decision Tree for the Two-Dimensional Vector Quantizer. the codebook and exploit this structure to reduce its size and/or improve the speed of the search. Tree-Structured VQ: An alternative to the previous method consists of structuring the codebook as a search tree. This strategy also allows a fast codebook search and, if desired, may provide additional functionality. The drawback of this approach is that, due to the structural constraints, the average distortion may be higher than that of a vector quantizer that uses an unstructured codebook. In a Tree-Structured Vector Quantizer, or TSVQ, the quantization is performed by traversing a tree and comparing the input vector V to the vectors associated with the nodes along a path. The final result of the quantization is the vector associated with a leaf node. Figure 7.123 shows an example of such quantizer. The input vector V is compared first to the vectors y1 and y2. The quatization continues on the branch corresponding to the vector having minimum distortion (the vector “closer” to V ). Assume that y2 is such a vector. In the next step V is compared to y5 and y6 and since y5 is closer, a comparison with C5 and C6 follows. The quantization ends at the leaf associated with C6 and the index corresponding to this vector is output and sent to the decoder.                Figure 7.123: Tree-Structured Vector Quantizer. Notice that in this example, the inner vectors are only used to speed up the search and that the decoder does not need to store them. On the other hand, vectors y2 and y5 encountered along the path can be considered a “coarser” quantization of V (with y2, in general, less precise than y5). By using this observation and choosing an indexing scheme that reflects the branching of the tree (such as the indexing used in a Huffman code), a decoder that stores the inner vectors can stop the quantization anywhere along 7.19 Vector Quantization 597 the path and output a lower quality quantization of V . This method is particularly useful when implementing a progressive compression scheme, where decoding a part of the bitstream results in a lower-quality representation of the compressed signal. Residual VQ: Progressive encoding can also be obtained with what is referred to as Residual Vector Quantizer. A residual vector quantizer consists of multiple cascaded stages, each quantizing the error (or residual) of the previous stage. Figure 7.124 illustrates such a quantizer.                        Figure 7.124: Four-Stages Residual Vector Quantizer. A residual vector quantizer simplifies the search by using several small codebooks instead of a single large one. Instead of quantizing a vector with, say, 16 bits with a codebook containing 216 = 65,536 codevectors, a four-stage residual quantizer could use, for example, four 4-bit indexes and four codebooks of 16 codevectors each (the codebooks of the quantizers Q1 through Q4 in the figure). Even with an exhaustive search, comparing an input vector to 64 codevectors still results in a practical algorithm. Each stage transmits a quantization index (I1 through I4 in the figure) that is independently dequantized at the decoder. To reconstruct the quantized vector, the decoder adds up the four codewords ˆ V , ˆE 1, ˆE 2, and ˆE3. Product VQ: If a vector V has too many components to be quantized as a whole, a simple strategy is to partition it into two or more smaller vectors and quantize the smaller vectors separately. This is clearly worse than quantizing the entire vector, since otherwise vector quantization wouldn’t be better than scalar quantization. However, if the sub-vectors are not strongly correlated, partitioning can result in an efficient and practical algorithm. The name Product Vector Quantizer derives from the fact that the codebook for the entire vector is the Cartesian product of the codebooks used to quantize the smaller vectors. An algorithm that partitions long vectors into sub-vectors of possibly different dimensions is described in Section 11.15.3. In the following we discuss an approach (often used in image coding) called Mean-Removed Vector Quantizer. Small pixel blocks in a digital image may repeat in the context of objects or areas having different luminosity. A vector quantizer would need different codewords to represent blocks that have the same pattern but different luminosity. For example, the vector representing a 4Χ4 image block containing a vertical edge will appear numerically different depending on whether the edge belongs to a well-lit object or to an object in a shaded area. 598 7. Image Compression A mean-removed vector quantizer addresses this problem by decomposing an ndimensional vector into a vector of dimension (n-1) and a scalar (Figure 7.125). As the name suggests, the scalar represents the mean value of the pixels in the vector and the (n - 1)-dimensional vector is obtained by subtracting the mean from each component and dropping a dimension. One of the dimensions can be discarded since the meanremoved vector has n-1 degrees of freedom (the sum of its components is always zero). The mean can now be quantized with a scalar quantizer and an (n - 1)-dimensional codebook can be used on the mean-removed vectors.                 Figure 7.125: Mean-Removed Vector Quantizer. The following illustrates the advantage of this approach. To quantize 4Χ4 image blocks with 16 bits, a traditional vector quantizer would require a codebook containing 65,536 vectors. A mean-removed quantizer would be much more practical since it could be designed with the mean quantized (very accurately) with seven bits, and the meanremoved vector with a codebook of only 512 entries. The approaches to the design of structured quantizers presented here are not mutually exclusive. The tree and the residual structures, for example, are often combined so that paths along the tree encode quantization residuals. Similarly, the vector quantizer used in a mean-removed VQ can have a residual structure, and so on. An in-depth discussion of these methods and the presentation of many more can be found in the excellent text by Gersho and Gray [Gersho and Gray 92]. 7.20 Adaptive Vector Quantization The basic vector quantization method of Section 7.19 uses either a fixed codebook or the LBG (or similar) algorithm to construct the codebook as it goes along. In all these cases the codebook consists of fixed-size entries, identical in size to the image block. The adaptive method described here, due to Constantinescu and Storer [Constantinescu and Storer 94a,b], uses variable-size image blocks and codebook entries. The codebook is called dictionary, because this method bears a slight resemblance to the various LZ methods. The method combines the single-pass, adaptive dictionary used by the various dictionary-based algorithms with the distance measures used by the different vector quantization methods to obtain good approximations of data blocks. At each step of the encoding process, the encoder selects an image block (a rectangle of any size), matches it to one of the dictionary entries, outputs a pointer to that entry, and updates the dictionary by adding one or more entries to it. The new entries are 7.20 Adaptive Vector Quantization 599 based on the current image block. The case of a full dictionary has to be taken care of, and is discussed below. Some authors refer to such a method as textual substitution, since it substitutes pointers for the original data. Extensive experimentation by the developers yielded compression ratios that equal or even surpass those of JPEG. At the same time, the method is fast (since it performs just one pass), simple (it starts with an empty dictionary, so no training is required), and reliable (the amount of loss is controlled easily and precisely by a single tolerance parameter). The decoder’s operation is similar to that of a vector quantization decoder, so it is fast and simple. The encoder has to select the image blocks carefully, making sure that they cover the entire image, that they cover it in a way that the decoder can mimic, and that there is not too much overlap between the different blocks. In order to select the blocks the encoder selects one or more growing points at each iteration, and uses one of them in the next iteration as the corner of a new image block. The block starts small and is increased in size (is “grown” from the corner) by the encoder as long as it can be matched with a dictionary entry. A growing point is denoted by G and is a triplet (x, y, q), where x and y are the coordinates of the point and q indicates the corner of the image block where the point is located. We assume that q values of 0, 1, 2, and 3 indicate the top-left, topright, bottom-left, and bottom-right corners, respectively (thus, q is a 2-bit number). An image block B anchored at a growing point G is denoted by B = (G,w, h), where w and h are the width and height of the block, respectively. We list the main steps of the encoder and decoder, then discuss each in detail. The encoder’s main steps are as follows: Step 1: Initialize the dictionary D to all the possible values of the image pixels. Initialize the pool of growing points (GPP) to one or more growing points using one of the algorithms discussed below. Step 2: Repeat Steps 3–7 until the GPP is empty. Step 3: Use a growing algorithm to select a growing point G from GPP. Step 4: Grow an image block B with G as its corner. Use a matching algorithm to match B, as it is being grown, to a dictionary entry with user-controlled fidelity. Step 5: Once B has reached the maximum size where it still can be matched with a dictionary entry d, output a pointer to d. The size of the pointer depends on the size (number of entries) of D. Step 6: Delete G from the GPP and use an algorithm to decide which new growing points (if any) to add to the GPP. Step 7: If D is full, use an algorithm to delete one or more entries. Use an algorithm to update the dictionary based on B. The operation of the encoder depends, therefore, on several algorithms. Each algorithm should be developed, implemented, and its performance tested with many test images. The decoder is much simpler and faster. Its main steps are as follows: Step 1: Initialize the dictionary D and the GPP as in Step 1 of the encoder. Step 2: Repeat Steps 3–5 until GPP is empty. Step 3: Use the encoder’s growing algorithm to select a growing point G from the GPP. 600 7. Image Compression Step 4: Input a pointer from the compressed stream, use it to retrieve a dictionary entry d, and place d at the location and position specified by G. Step 5: Update D and the GPP as in Steps 6–7 of the encoder. The remainder of this section discusses the various algorithms needed, and shows an example. Algorithm: Select a growing point G from the GPP. The simplest methods are LIFO (which probably does not yield good results) and FIFO (which is used in the example below), but the coverage methods described here may be better, since they determine how the image will be covered with blocks. Wave Coverage: This algorithm selects from among all the growing points in the GPP the point G = (xs, ys, 0) that satisfies xs + ys . x + y, for any G(x, y, 0) in GPP (notice that the growing point selected has q = 0, so it is located at the upper-left corner of an image block). When this algorithm is used, it makes sense to initialize the GPP with just one point, the top-left corner of the image. If this is done, then the wave coverage algorithm will select image blocks that cover the image in a wave that moves from the upper-left to the lower-right corners of the image (Figure 7.126a). Circular Coverage: Select the growing point G whose distance from the center of the image is minimal among all the growing points G in the GPP. This covers the image with blocks located in a growing circular region around the center (Figure 7.126b). Diagonal Coverage: This algorithm selects from among all the growing points in the GPP the point G = (xs, ys, q) that satisfies |xs - ys| . |x - y|, for any G(x, y, p) in GPP (notice that the growing point selected may have any q). The result will be blocks that start around the main diagonal and move away from it in two waves that run parallel to it. Algorithm: Matching an image block B (with a growing point G as its corner) to a dictionary entry. It seems that the best approach is to start with the smallest block B (just a single pixel) and try to match bigger and bigger blocks to the dictionary entries, until the next increase of B does not find any dictionary entry to match B to the tolerance specified by the user. The following parameters control the matching: 1. The distance measure. Any of the measures proposed in Equation (7.29) can be used. 2. The tolerance. A user-defined real parameter t that’s the maximum allowed distance between an image block B and a dictionary entry. 3. The type of coverage. Since image blocks have different sizes and grow from the growing point in different directions, they can overlap (Figure 7.126a). Imagine that an image block B has been matched, but it partly covers some other blocks. We can compute the distance using just those parts of B that don’t overlap any other blocks. This means that the first block that covers a certain image area will determine the quality of the decompressed image in that area. This can be called first coverage. The opposite is last coverage, where the distance between an image block B and a dictionary entry is computed using all of B (except those parts that lie outside the image), regardless of any 7.20 Adaptive Vector Quantization 601 overlaps. It is also possible to have average coverage, where the distance is computed for all of B, but in a region of overlap, any pixel value used to calculate the distance will be the average of pixel values from all the image blocks in the region. 4. The elementary subblock size l. This is a “normalization” parameter that compensates for the different block sizes. We know that image blocks can have different sizes. It may happen that a large block will match a certain dictionary entry to within the tolerance, while some regions within the block will match poorly (and other regions will yield a very good match). An image block should therefore be divided into subblocks of size lΧl (where the value of l depends on the image size, but is typically 4 or 5), and each dictionary entry should similarly be divided. The matching algorithm should calculate the distance between each image subblock and the corresponding dictionary entry subblock. The maximum of these distances should be selected as the distance between the image block and the dictionary entry. (a) (b) (c) Figure 7.126: (a) Wave Coverage. (b) Growing Points. (c) Dictionary Update. Algorithm: GPP update. A good policy is to choose as new growing points those points located at or near the borders of the partially encoded image (Figure 7.126b). This causes new image blocks to be adjacent to old ones. This policy makes sense because an old block has contributed to new dictionary entries, and those entries are perfect candidates to match a new, adjacent block, because adjacent blocks generally don’t differ much. Initially, when the partially encoded image is small, consisting mainly of single-pixel blocks, this policy adds two growing points to the GPP for each small block added to the encoded image. Those are the points below and to the right of the image block (see example below). Another good location for new growing points is any new concave corners that have been generated by adding the recent image block to the partially encoded image (Figure 7.126a,b). We show below that when a new image block is grown from such a corner point, it contributes two or three new entries to the dictionary. Algorithm: Dictionary update. This is based on two principles: (1) Each matched block added to the image-so-far should be used to add one or more dictionary entries; (2) the new entries added should contain pixels in the image-encoded-so-far (so the decoder can update its dictionary in the same way). Denoting the currently matched block by B, a simple policy is to add to the dictionary the two blocks obtained by adding one column of pixels to the left of B and one row of pixels above it. Sometimes, as Figure 7.126c 602 7. Image Compression shows, the row or column can be added to just part of B. Later experiments with the method added a third block to the dictionary, the block obtained by adding the column to the left of B and the row above it. Since image blocks B start small and square, this third dictionary entry is also square. Adding this third block improves compression performance significantly. Algorithm: Dictionary deletion. The dictionary keeps growing but it can grow only so much. When it reaches its maximum size, we can select from among the following ideas: (1) Delete the least recently used (LRU) entry (except that the original entries, all the possible pixel values, should not be deleted); (2) freeze the dictionary (i.e., don’t add any new entries); and (3) delete the entire dictionary (except the original entries) and start afresh.  Exercise 7.28: Suggest another approach to dictionary deletion. (Hint: See Section 6.14.) Experiments performed by the method’s developers suggest that the basic algorithm is robust and does not depend much on the particular choice of the algorithms above, except for two choices that seem to improve compression performance significantly: 1. Wave coverage seems to offer better performance than circular or diagonal coverage. 2. Visual inspection of many decompressed images shows that the loss of data is more visible to the eye in image regions that are smooth (i.e., uniform or close to uniform). The method can therefore be improved by using a smaller tolerance parameter for such regions. A simple measure A of smoothness is the ratio V/M, where V is the variance of the pixels in the region (see page 624 for the definition of variance) and M is their mean. The larger A, the more active (i.e., the less smooth) is the region. The tolerance t is determined by the user, and the developers of the method suggest the use of smaller tolerance values for smooth regions as follows: if5A . 0.05, use 0.4t, 0.05 < A . 0.1, use 0.6t, A > 0.1, use t. Example: We select an image that consists of 4-bit pixels. The nine pixels at the top-left corner of this image are shown in Figure 7.127a. They have image coordinates ranging from (0, 0) to (2, 2). The first 16 entries of the dictionary D (locations 0 through 15) are initialized to the 16 possible values of the pixels. The GPP is initialized to the top-left corner of the image, position (0, 0). We assume that the GPP is used as a (FIFO) queue. Here are the details of the first few steps. Step 1: Point (0, 0) is popped out of the GPP. The pixel value at this position is 8. Since the dictionary contains just individual pixels, not blocks, this point can be matched only with the dictionary entry containing 8. This entry happens to be located in dictionary location 8, so the encoder outputs the pointer 8. This match does not have any concave corners, so we push the point on the right of the matched block, (0, 1), and the point below it, (1, 0), into the GPP. Step 2: Point (1, 0) is popped out of the GPP. The pixel value at this position is 4. The dictionary still contains just individual pixels, not any bigger blocks, so this point can be matched only with the dictionary entry containing 4. This entry is located in 7.21 Block Truncation Coding 603 dictionary location 4, so the encoder outputs the pointer 4. The match does not have any concave corners, so we push the point on the right of the matched block, (1, 1), and the point below it, (2, 0), into the GPP. The GPP now contains (from most recent to least recent) points (1, 1), (2, 0), and (0, 1). The dictionary is updated by appending to it (at location 16) the block 84 as shown in Figure 7.127b. 012... 0 1 2 . . . 8 2 3 4 5 6 7 8 9 0,1,2,. . . ,15, 84 (b) 0,1,2,. . . ,15, 84 , 8 2 , 47 , 25 , 4 5 . (a) (c) Figure 7.127: An Image and a Dictionary. Step 3: Point (0, 1) is popped out of the GPP. The pixel value at this position is 2. The dictionary contains the 16 individual pixel values and one bigger block. The best match for this point is with the dictionary entry containing 2, which is at dictionary location 2, so the encoder outputs the pointer 2. The match does not have any concave corners, so we push the point on the right of the matched block, (0, 2), and the point below it, (1, 1), into the GPP. The GPP is now (from most recent to least recent) points (0, 2), (1, 1), and (2, 0). The dictionary is updated by appending to it (at location 17) the block 8 2 (Figure 7.127c).  Exercise 7.29: Do the next two steps. 7.21 Block Truncation Coding Quantization is an important technique for data compression in general and for image compression in particular. Any quantization method should be based on a principle that determines what data items to quantize and by how much. The principle used by the block truncation coding (BTC) method and its variants is to quantize pixels in an image while preserving the first two or three statistical moments. The main reference is [Dasarathy 95], which includes an introduction and copies of the main papers in this area. In the basic BTC method, the image is divided into blocks (normally 4Χ4 or 8Χ8 pixels each). Assuming that a block contains n pixels with intensities p1 through pn, the first two moments are the mean and variance, defined as ―p = 1 n n i=1 pi, (7.31) 604 7. Image Compression and p2 = 1 n n i=1 p2i , (7.32) respectively. The standard deviation of the block is . = )p2 - ―p. (7.33) The principle of the quantization is to select three values, a threshold pthr, a high value p+, and a low value p-. Each pixel is replaced by either p+ or p-, such that the first two moments of the new pixels (i.e., their mean and variance) will be identical to the original moments of the pixel block. The rule of quantization is that a pixel pi is quantized to p+ if it is greater than the threshold, and is quantized to p- if it is less than the threshold (if pi equals the threshold, it can be quantized to either value). Thus, pi ‹ p+, if pi . pthr, p-, if pi < pthr. Intuitively, it is clear that the mean ―p is a good choice for the threshold. The high and low values can be determined by writing equations that preserve the first two moments, and solving them. We denote by n+ the number of pixels in the current block that are greater than or equal to the threshold. Similarly, n- stands for the number of pixels that are smaller than the threshold. The sum n+ +n- equals, of course, the number of pixels n in the block. Once the mean ―p has been computed, both n+ and n- are easy to calculate. Preserving the first two moments is expressed by the two equations n―p = n-p- - n+p+, np2 = n-(p-)2 - n+(p+)2. (7.34) These are easy to solve even though they are nonlinear, and the solutions are p- = ―p - .4n+ n- , p+ = ―p + .4n- n+ . (7.35) These solutions are generally real numbers, but they have to be rounded to the nearest integer, which implies that the mean and variance of the quantized block may be somewhat different from those of the original block. Notice that the solutions are located on the two sides of the mean ―p at distances that are proportional to the standard deviation . of the pixel block. Example: We select the 4Χ4 block of 8-bit pixels ... 121 114 56 47 37 200 247 255 16 0 12 169 43 5 7 251 ... . The mean is ―p = 98.75, so we count n+ = 7 pixels greater than the mean and n- = 16 - 7 = 9 pixels less than the mean. The standard deviation is . = 92.95, so the high 7.21 Block Truncation Coding 605 and low values are p+ = 98.75 + 92.9549 7 = 204.14, p- = 98.75 - 92.9547 9 = 16.78. They are rounded to 204 and 17, respectively. The resulting block is ... 204 204 17 17 17 204 204 204 17 17 17 204 17 17 17 204 ... , and it is clear that the original 4Χ4 block can be compressed to the 16 bits ... 1 1 0 0 0 1 1 1 0 0 0 1 0 0 0 1 ... , plus the two 8-bit values 204 and 17; a total of 16 + 2Χ8 bits, compared to the 16Χ8 bits of the original block. The compression factor is 16Χ8 16 + 2Χ8 = 4.  Exercise 7.30: Do the same for the 4Χ4 block of 8-bit pixels ... 136 27 144 216 172 83 43 219 200 254 1 128 64 32 96 25 ... . It is clear that the compression factor of this method is known in advance, because it depends on the block size n and on the number of bits b per pixel. In general, the compression factor is bn n + 2b . Figure 7.128 shows the compression factors for b values of 2, 4, 8, 12, and 16 bits per pixel, and for block sizes n between 2 and 16. The basic BTC method is simple and fast. Its main drawback is the way it loses image information, which is based on pixel intensities in each block, and not on any properties of the human visual system. Because of this, images compressed under BTC tend to have a blocky character when decompressed and reconstructed. This led many researchers to develop enhanced and extended versions of the basic BTC, some of which are briefly mentioned here. 1. Encode the results before writing them on the compressed stream. In basic BTC, each original block of n pixels becomes a block of n bits plus two numbers. It is possible to 606 7. Image Compression                        Figure 7.128: BTC Compression Factors for Two Preserved Moments. accumulate B blocks of n bits each, and encode the Bn bits using run-length encoding (RLE). This should be followed by 2B values, and those can be encoded with prefix codes. 2. Sort the n pixels pi of a block by increasing intensity values, and denote the sorted pixels by si. The first n- pixels si have values that are less than the threshold, and the remaining n+ pixels si have larger values. The high and low values, p+ and p-, are now the values that minimize the square-error expression n--1 i=1 si - p- 2 + n+ i=n- si - p+ 2 . (7.36) These values are p- = 1 n- n--1 i=1 si and p+ = 1 n+ n+ i=n- si. The threshold in this case is determined by searching all n-1 possible threshold values and choosing the one that produces the lowest value for the square-error expression of Equation (7.36). 3. Sort the pixels as in 2 above, and change Equation (7.36) to a similar one using absolute values instead of squares: n--1 i=1 |si - p-| + n+ i=n- |si - p+|. (7.37) The high and low values p+ and p- should be, in this case, the medians of the low and high intensity sets of pixels, respectively. Thus, p- should be the median of the set 7.21 Block Truncation Coding 607 {si|i = 1, 2, . . . , n- - 1}, and p+ should be the median of the set {si|i = n-, . . . , n+}. The threshold value is determined as in 2 above. 4. This is an extension of the basic BTC method, where three moments are preserved, instead of two. The three equations that result make it possible to calculate the value of the threshold, in addition to those of the high (p+) and low (p-) parameters. The third moment is p3 = 1 n n i=1 p3i , (7.38) and its preservation yields the equation np3 = n- p- 3 - n+ p+ 3 . (7.39) Solving the three equations (7.34) and (7.39) with p-, p+, and n+ as the unknowns yields n+ = n 2 1 + . ..2 + 4 , where . = 3―p(p2) - p3 - 2(―p)3 .3 . And the threshold pthr is selected as pixel pn+. 5. BTC is a lossy compression method where the data being lost is based on the mean and variance of the pixels in each block. This loss has nothing to do with the human visual system [Salomon 99], or with any general features of the image being compressed, so it may cause artifacts and unrecognizable regions in the decompressed image. One especially annoying feature of the basic BTC is that straight edges in the original image become jagged in the reconstructed image. The edge following algorithm [Ronson and Dewitte 82] is an extension of the basic BTC, where this tendency is minimized by identifying pixels that are on an edge, and using an additional quantization level for blocks containing such pixels. Each pixel in such a block is quantized into one of three different values instead of the usual two. This reduces the compression factor but improves the visual appearance of the reconstructed image. Another extension of the basic BTC is a three-stage process, proposed in [Dewitte and Ronson 83] where the first stage is basic BTC, resulting in a binary matrix and two numbers. The second stage classifies the binary matrix into one of three categories as follows: a. The block has no edges (indicating a uniform or near-uniform region). b. There is a single edge across the block. c. There is a pair of edges across the block. The classification is done by counting the number of transitions from 0 to 1 and from 1 to 0 along the four edges of the binary matrix. Blocks with zero transitions or with more than four transitions are considered category a (no edges). Blocks with two transitions are considered category b (a single edge), and blocks with four transitions are classified as category c (two edges). 608 7. Image Compression Stage three matches the classified block of pixels to a particular model depending on its classification. Experiments with this method suggest that about 80% of the blocks belong in category a, 12% are category b, and 8% belong in category c. The fact that category a blocks are so common implies that the basic BTC can be used in most cases, and the particular models used for categories b and c improve the appearance of the reconstructed image without significantly degrading the compression time and compression factor. 6. The original BTC is based on the principle of preserving the first two statistical moments of a pixel block. The variant described here changes this principle to preserving the first two absolute moments. The first absolute moment is given by pa = 1 n n i=1 |pi - ―p|, and it can be shown that this implies p+ = ―p + npa 2n+ and p- = ―p - npa 2n- , which shows that the high and low values are simpler than those used by the basic BTC. They are obtained from the mean by adding and subtracting quantities that are proportional to the absolute first moment pa and inversely proportional to the numbers n+ and n- of pixels on both sides of this mean. Moreover, these high and low values can also be written as p+ = 1 n+ pi.―p pi and p- = 1 n- pi<―p pi, simplifying their computation even more. In addition to these improvements and extensions there are BTC variants that deviate significantly from the basic idea. The three-level BTC is a good example of the latter. It quantizes each pixel pi into one of three values, instead of the original two. This, of course, requires two bits per pixel, thereby reducing the compression factor. The method starts by scanning all the pixels in the block and finding the range . of intensities . = max(pi) - min(pi), for i = 1, 2, . . . , n. Two mean values, high and low, are defined by ph = max(pi) - ./3 and pl = min(pi)+./3. The three quantization levels, high, low, and mid, are now given by p+ = 12[ph + max(pi)], p- = 12 [pl + min(pi)], pm = 12 [pl + ph] . A pixel pi is compared to p+, p-, and pm and is quantized to the nearest of these values. It takes two bits to express the result of quantizing pi, so the original block of n b-bit 7.22 Context-Based Methods 609 pixels becomes a block of 2n bits. In addition, the values of max(pi) and min(pi) have to be written on the compressed stream as two b-bit numbers. The resulting compression factor is bn 2n + 2b , lower than in basic BTC. Figure 7.129 shows the compression factors for b = 4, 8, 12, and 16 and for n values in the range [2, 16]. 7 5 3 1 2 4 b=4 b=8 b=12 b=16 6 8 Block size Compression factor 10 12 14 16 Figure 7.129: BTC Compression Factors for Three Quantization Levels. In spite of worse compression, the three-level BTC has an overall better performance than the basic BTC because it requires simpler calculations (in particular, the second moment is not needed) and because its reconstructed images are of higher quality and don’t have the blocky appearance typical of images compressed and decompressed under basic BTC. 7.22 Context-Based Methods Most image-compression methods are based on the observation that for any randomlyselected pixel in the image, its near neighbors tend to have the same value as the pixel (or similar values). A context-based image compression method generalizes this observation. It is based on the idea that the context of a pixel can be used to predict (estimate the probability of) the pixel. Such a method compresses an image by scanning it pixel by pixel, examining the context of every pixel, and assigning it a probability depending on how many times the same context was seen in the past. The pixel and its assigned probability are then sent to an arithmetic encoder that does the actual encoding. The methods described here are due to Langdon [Langdon and Rissanen 81] and Moffat [Moffat 91], and apply to monochromatic (bi-level) images. The context of a pixel consists of some of its neighbor pixels that have already been processed. The diagram below shows a possible sevenpixel context (the pixels marked P) made up of five pixels above and two on the left of 610 7. Image Compression the current pixel X. The pixels marked “?” haven’t been input yet, so they cannot be included in the context. · · P P P P P · · · · P P X ? ? ? ? The main idea is to use the values of the seven context pixels as a 7-bit index to a frequency table, and find the number of times a 0 pixel and a 1 pixel were seen in the past with the same context. Here is an example: · · 1 0 0 1 1 · · · · 0 1 X ? ? ? ? Since 10011012 = 77, location 77 of the frequency table is examined. We assume that it contains a count of 15 for a 0 pixel and a count of 11 for a 1 pixel. The current pixel is therefore assigned probability 15/(15 + 11) . 0.58 if it is 0 and 11/26 . 0.42 if it is 1. 77 · · · 15 · · · · · · 11 · · · One of the counts at location 77 is then incremented, depending on what the current pixel is. Figure 7.130 shows ten possible ways to select contexts. They range from a 1-bit, to a 22-bit context. In the latter case, there may be 222 . 4.2 million contexts, most of which will rarely, if ever, occur in the image. Instead of maintaining a huge, mostly empty array, the frequency table can be implemented as a binary search tree or a hash table. For short, 7-bit, contexts, the frequency table can be an array, resulting in fast search. P X ? ? ? P P X ? ? ? P P P P X ? ? ? · P P P P P P X ? ? ? P P P P P P P P X ? ? ? p= 1 2 4 6 8 P P P P P P P P P P X ? ? ? P P P P P P P P P P P P X ? ? ? P P P P P P P P P P P P P P X ? ? ? P P P P P P P P P P P P P P P P P P X ? ? ? P P P P P P P P P P P P P P P P P P P P P P X ? ? ? p = 10 12 14 18 22 Figure 7.130: Context Pixel Patterns. “Bless me, bless me, Sept,” returned the old lady, “you don’t see the context! Give it back to me, my dear.” —Charles Dickens, The Mystery of Edwin Drood (1870) Experience shows that longer contexts result in better compression, up to about 12–14 bits. Contexts longer than that result in worse compression, indicating that a 7.22 Context-Based Methods 611 pixel in a typical image does not “relate” to distant pixels (correlations between pixels are typically limited to a distance of 1–2). As usual for a context-based compression method, care should be taken to avoid zero probabilities. The frequency table should thus be initialized to nonzero values, and experience shows that the precise way of doing this does not affect the compression performance significantly. It seems best to initialize every table entry to 1. When the process starts, the first pixels to be scanned don’t have any neighbors above them or to the left. A simple way to handle this is to assume that any nonexistent neighbors are zeros. It is as if the image was extended by adding as many rows of zero pixels as necessary on top and as many zero columns on the left. The 2-bit context of Figure 7.130 is now used, as an example, to encode the first row of the 4 Χ 4 image 0 0 0 0 0 1 0 1 1 0 0 1 0 1 0 1 0 0 1 0 0 0 1 0 The results are summarized in Table 7.131. Number Pixel Context Counts Probability New counts 1 1 00=0 1,1 1/(1 + 1) = 1/2 1, 2 2 0 01=1 1,1 1/2 2, 1 3 1 00=0 1,2 2/3 1, 3 4 1 01=1 2,1 1/3 2, 2 Table 7.131: Counts and Probabilities for First Four Pixels.  Exercise 7.31: Continue Table 7.131 for the next row of four pixels 0101. The contexts of Figure 7.130 are not symmetric about the current pixel, since they must use pixels that have already been processed (“past” pixels). If the algorithm scans the image by rows, those will be the pixels above the current pixel or to its left. In practical work it is impossible to include “future” pixels in the context (because the decoder wouldn’t have their values when decoding the current pixel), but for experiments, where there is no need to actually decode the compressed image, it is possible to store the entire image in memory so that the encoder can examine any pixel at any time. Experiments performed with symmetric contexts have shown that compression performance can improve by as much as 30%. (The MLP method, Section 7.25, provides an interesting twist to the question of a symmetric context.)  Exercise 7.32: (Easy.) Why is it possible to use “future” pixels in an experiment but not in practice? It would seem that the image, or part of it, could be stored in memory and the encoder could use any pixel as part of a context? 612 7. Image Compression One disadvantage of a large context is that it takes the algorithm longer to “learn” it. A 20-bit context, for example, allows for about a million different contexts. It takes many millions of pixels to collect enough counts for all those contexts, which is one reason large contexts do not result in better compression. One way to improve our method is to implement a two-level algorithm that uses a long context only if that context has already been seen Q times or more (where Q is a parameter, typically set to a small value such as 2 or 3). If a context has been seen fewer than Q times, it is deemed unreliable, and only a small subset of it is used to predict the current pixel. Figure 7.132 shows four such contexts, where the pixels of the subset are labeled S. The notation p, q means a two-level context of p bits with a subset of q bits. P P P P P S S S S P S S X ? ? ? P P P S P P S S S S S P S S X ? ? ? P P P P S S S P P S S S S S P P S S X ? ? ? P P P P P P P S S S P P P S S S S S P P S S X ? ? ? p, q = 12,6 14,8 18,10 22,10 Figure 7.132: Two-Level Context Pixel Patterns. Experience shows that the 18, 10 and 22, 10 contexts result in better, although not revolutionary, compression. 7.23 FELICS FELICS is an acronym for Fast, Efficient, Lossless Image Compression System. It is a special-purpose compression method [Howard and Vitter 93] designed for grayscale images and it competes with the lossless mode of JPEG (Section 7.10.5). It is fast and it generally produces good compression. However, it cannot compress an image to below one bit per pixel, so it is not a good choice for bi-level or for highly redundant images. The principle of FELICS is to code each pixel with a variable-length code based on the values of two of its previously seen neighbor pixels. Figure 7.133a shows the two known neighbors A and B of some pixels P. For a general pixel, these are the neighbors above it and to its left. For a pixel in the top row, these are its two left neighbors (except for the first two pixels of the image). For a pixel in the leftmost column, these are the first two pixels of the line above it. Notice that the first two pixels of the image don’t have any previously seen neighbors, but since there are only two of them, they can be output without any encoding, causing just a slight degradation in the overall compression.  Exercise 7.33: What model is used by FELICS to predict the current pixel? Consider the two neighbors A and B of a pixel P. We use A, B, and P to denote both the three pixels and their intensities (grayscale values). We denote by L and H the neighbors with the smaller and the larger intensities, respectively. Pixel P should be assigned a variable-length code depending on where the intensity P is located relative to L and H. There are three cases: 7.23 FELICS 613 A B P A B P B A P 10... 0... 11... L H Probability (a) (b) Figure 7.133: (a) The Two Neighbors. (b) The Three Regions. 1. The intensity of pixel P is between L and H (it is located in the central region of Figure 7.133b). This case is known experimentally to occur in about half the pixels, and P is assigned, in this case, a code that starts with 0. (A special case occurs when L = H. In such a case, the range [L, H] consists of one value only, and the chance that P will have that value is small.) The probability that P will be in this central region is almost, but not completely, flat, so P should be assigned a binary code that has about the same size in the entire region but is slightly shorter at the center of the region. 2. The intensity of P is lower than L (P is in the left region). The code assigned to P in this case starts with 10. 3. P’s intensity is greater than H. P is assigned a code that starts with 11. When pixel P is in one of the outer regions, the probability that its intensity will differ from L or H by much is small, so P can be assigned a long code in these cases. The code assigned to P should therefore depend heavily on whether P is in the central region or in one of the outer regions. Here is how the code is assigned when P is in the central region. We need H-L+1 variable-length codes that will not differ much in size and will, of course, satisfy the prefix property. We set k = log2(H -L+1) and compute integers a and b by a = 2k+1 - (H - L + 1), b= 2(H - L + 1 - 2k). Example: If H - L = 9, then k = 3, a = 23+1 - (9 + 1) = 6, and b = 2(9 + 1 - 23) = 4. We now select the a codes 2k -1, 2k -2,. . . expressed as k-bit numbers, and the b codes 0, 1, 2, . . . expressed as (k + 1)-bit numbers. In the example above, the a codes are 8 - 1 = 111, 8 - 2 = 110, through 8 - 6 = 010, and the b codes, 0000, 0001, 0010, and 0011. The a short codes are assigned to values of P in the middle of the central region, and the b long codes are assigned to values of P closer to L or H. Notice that b is even, so the b codes can always be partitioned into two equal sets. Table 7.134 shows how ten such codes can be assigned in the case L = 15, H = 24. When P is in one of the outer regions, say the upper one, the value P-H should be assigned a variable-length code whose size can grow quickly as P - H gets bigger. One way to do this is to select a small nonnegative integer m (typically 0, 1, 2, or 3) and to assign the integer n a two-part code. The second part is the lower m bits of n, and 614 7. Image Compression Pixel Region Pixel P code code L=15 0 0000 16 0 0010 17 0 010 18 0 011 19 0 100 20 0 101 21 0 110 22 0 111 23 0 0001 H=24 0 0011 Table 7.134: The Codes for the Central Region. the first part is the unary code of [n without its lower m bits] (see Section 3.1 for this code). Example: If m = 2, then n = 11012 is assigned the code 110|01, since 110 is the unary code of 11. This code is a special case of the Golomb code (Section 3.24), where the parameter b is a power of 2 (2m). Table 7.135 shows some examples of this code for m = 0, 1, 2, 3 and n = 1, 2, . . . , 9. The value of m used in any particular compression job can be selected, as a parameter, by the user. Pixel Region m = P P–H code 0 1 2 3 H+1=25 1 11 0 00 000 0000 26 2 11 10 01 001 0001 27 3 11 110 100 010 0010 28 4 11 1110 101 011 0011 29 5 11 11110 1100 1000 0100 30 6 11 111110 1101 1001 0101 31 7 11 1111110 11100 1010 0110 32 8 11 11111110 11101 1011 0111 33 9 11 111111110 111100 11000 10000 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Table 7.135: The Codes for an Outer Region.  Exercise 7.34: Given the 4Χ4 bitmap of Figure 7.136, calculate the FELICS codes for the three pixels with values 8, 7, and 0. 7.24 Progressive FELICS 615 2 5 7 12 3 0 11 10 2 1 8 15 4 13 11 9 Figure 7.136: A 4Χ4 Bitmap. 7.24 Progressive FELICS The original FELICS method can easily be extended to progressive compression of images because of its main principle. FELICS scans the image in raster order (row by row) and encodes a pixel based on the values of two of its (previously seen and encoded) neighbors. Progressive FELICS [Howard and Vitter 94a] works similarly, but it scans the pixels in levels. Each level uses the k pixels encoded in all previous levels to encode k more pixels, so the number of encoded pixels doubles after each level. Assuming that the image consists of nΧn pixels, and the first level starts with just four pixels, consecutive levels result in 4, 8, . . . , n2 8 , n2 4 , n2 2 , n2 pixels. Thus, the number of levels is the number of terms, 2 log2 n-1, in this sequence. (This property is easy to prove. The first term can be written 4 = n2 2-2n2 = n2 22 log n-2 . Terms in the sequence therefore contain powers of 2 that go from 0 to 2 log2 n - 2, showing that there are 2 log2 n - 1 terms.) Figure 7.137 shows the pixels encoded in most of the levels of a 16Χ16-pixel image. Figure 7.138 shows how the pixels of each level are selected. In Figure 7.138a there are 8 Χ 8 = 64 pixels, one quarter of the final number, arranged in a square grid. Each group of four pixels is used to encode a new pixel, so Figure 7.138b has 128 pixels, half the final number. The image of Figure 7.138b is then rotated 45. and scaled by factors of .2 . 1.414 in both directions, to produce Figure 7.138c, which is a square grid that looks exactly like Figure 7.138a. The next step (not shown in the figure) is to use every group of 4Χ4 pixels in Figure 7.138c to encode a pixel, thereby encoding the remaining 128 pixels. In practice, there is no need to actually rotate and scale the image; the program simply alternates between xy and diagonal coordinates. Each group of four pixels is used to encode the pixel at its center. Notice that in early levels the four pixels of a group are far from each other and are therefore not correlated, thereby resulting in poor compression. However, the last two levels encode 3/4 of the total number of pixels, and these levels contain compact groups. Two of the four pixels of a group are selected to encode the center pixel, and are designated L and H. Experience shows that the best choice for L and H is the two median pixels (page 554), the ones with the middle values (i.e., not the maximum or the minimum pixels of the group). Ties can be resolved in any way, but it should be consistent. If the two medians in a group are the same, then the median and the minimum (or the median and the maximum) pixels can be selected. The two selected pixels, L and H, are 616 7. Image Compression Figure 7.137: Some Levels of a 16Χ16 Image. 7.24 Progressive FELICS 617 used to encode the center pixel in the same way FELICS uses two neighbors to encode a pixel. The only difference is that a new prefix code (Section 7.24.1) is used, instead of the Golomb code.  Exercise 7.35: Why is it important to resolve ties in a consistent way? 7.24.1 Subexponential Code In early levels, the four pixels used to encode a pixel are far from each other. As more levels are progressively encoded the groups get more compact, so their pixels get closer. The encoder uses the absolute difference between the L and H pixels in a group (the context of the group) to encode the pixel at the center of the group, but a given absolute difference means more for late levels than for early ones, because the groups of late levels are smaller, so their pixels are more correlated. The encoder should therefore scale the difference by a weight parameter s that gets heavier from level to level. The specific value of s is not critical, and experiments recommend the value 12. The prefix code used by progressive FELICS is called subexponential. They are related to the Rice codes of Section 10.9. Like the Golomb code (Section 3.24), this new code depends on a parameter k . 0. The main feature of the subexponential code is its length. For integers n < 2k+1 the code length increases linearly with n, but for larger n, it increases logarithmically. The subexponential code of the nonnegative integer n is computed in two steps. In the first step, values b and u are calculated by b = k if n < 2k, log2 n if n . 2k, and u = 0 ifn < 2k, b - k + 1 if n . 2k. In the second step, the unary code of u (in u+1 bits), followed by the b least-significant bits of n, becomes the subexponential code of n. Thus, the total size of the code is u+1+b = k + 1 if n < 2k, 2log2 n - k + 2 if n . 2k. Table 7.139 shows examples of the subexponential code for various values of n and k. It can be shown that for a given n, the code lengths for consecutive values of k differ by at most 1. If the value of the pixel to be encoded lies between those of L and H, the pixel is encoded as in FELICS. If it lies outside the range [L,H], the pixel is encoded by using the subexponential code where the value of k is selected by the following rule: Suppose that the current pixel P to be encoded has context C. The encoder maintains a cumulative total, for some reasonable values of k, of the code length the encoder would have if it had used that value of k to encode all pixels encountered so far in context C. The encoder then uses the k value with the smallest cumulative code length to encode pixel P. 618 7. Image Compression (a) (b) (c) Code more pixels Rotate and scale Figure 7.138: Rotation and Scaling. 7.25 MLP 619 n k= 0 k = 1 k = 2 k = 3 k = 4 k = 5 0 0| 0|0 0|00 0|000 0|0000 0|00000 1 10| 0|1 0|01 0|001 0|0001 0|00001 2 110|0 10| 0 0|10 0|010 0|0010 0|00010 3 110|1 10| 1 0|11 0|011 0|0011 0|00011 4 1110|00 110|00 10|00 0|100 0|0100 0|00100 5 1110|01 110|01 10|01 0|101 0|0101 0|00101 6 1110|10 110|10 10|10 0|110 0|0110 0|00110 7 1110|11 110|11 10|11 0|111 0|0111 0|00111 8 11110|000 1110|000 110|000 10|000 0|1000 0|01000 9 11110|001 1110|001 110|001 10|001 0|1001 0|01001 10 11110|010 1110|010 110|010 10|010 0|1010 0|01010 11 11110|011 1110|011 110|011 10|011 0|1011 0|01011 12 11110|100 1110|100 110|100 10|100 0|1100 0|01100 13 11110|101 1110|101 110|101 10|101 0|1101 0|01101 14 11110|110 1110|110 110|110 10|110 0|1110 0|01110 15 11110|111 1110|111 110|111 10|111 0|1111 0|01111 16 111110|0000 11110|0000 1110|0000 110|0000 10|0000 0|10000 Table 7.139: Some Subexponential Codes. 7.25 MLP Note. The MLP image compression method described in this section is different from and unrelated to the MLP (meridian lossless packing) audio compression method of Section 10.7. The identical acronyms are an unfortunate coincidence. Text compression methods can use context to predict (i.e., to estimate the probability of) the next character of text. Context can also be used to predict the intensity of the next pixel in image compression, but this is more complex for two reasons: (1) An image is two-dimensional, allowing for many possible contexts, and (2) a digital image is often the result of digitizing an analog image. The intensity of any individual pixel is therefore determined by the details of scanning and may differ from the “ideal” intensity. The multilevel progressive method (MLP) described here [Howard and Vitter 92a], is a computationally intensive, lossless method for compressing grayscale images. It uses context to predict the intensities of pixels, then uses arithmetic coding to encode the difference between the prediction and the actual value of a pixel (the error). The Laplace distribution is used to estimate the probability of the error. The method combines four separate steps: (1) pixel sequencing, (2) prediction (image modeling), (3) error modeling (by means of the Laplace distribution), and (4) arithmetically encoding the errors. MLP is also progressive, encoding the image in levels, where the pixels of each level are selected as in progressive FELICS. When the image is decoded, each level adds details to the entire image, not just to certain parts, so a user can view the image as it is being decoded and decide in real time whether to accept or reject it. This feature is useful when an image has to be selected from a large archive of compressed images. The user can browse through images very fast, without having to wait for any image to be completely decoded and displayed. Another advantage of progressive compression is 620 7. Image Compression that it provides a natural lossy option. The encoder may be told to stop encoding before it reaches the last level (thereby encoding only half the total number of pixels) or before it reaches the next to last level (encoding only a quarter of the total number of pixels). Such an option results in excellent compression but a loss of image data. The decoder may be told to use interpolation to determine the intensities of any missing pixels. Like any compression method for grayscale images, MLP can be used to compress color images. The original color image should be separated into three color components, and each component compressed individually as a grayscale image. Following is a detailed description of the individual MLP encoding steps. What we know is not much. What we do not know is immense. —(Allegedly Laplace’s last words.) 7.25.1 Pixel Sequencing Pixels are selected in levels, as in progressive FELICS, where each level encodes the same number of pixels as all the preceding levels combined, thereby doubling the number of encoded pixels. This means that the last level encodes half the number of pixels, the level preceding it encodes a quarter of the total number, and so on. The first level should start with at least four pixels, but may also start with 8, 16, or any desired power of 2. . . . . . . • . . . • • . . • • • . • • ? • • . • • • . . • • . . . • . . . . . . . . . . . . . . . . . . . . . . . . . . . . • . . . . • . . . •? • . . . • • . . (a) (b) Figure 7.140: (a) Sixteen Neighbors. (b) Six Neighbors. Man follows only phantoms. —(Allegedly Laplace’s last words.) 7.25 MLP 621 7.25.2 Prediction A pixel is predicted by calculating a weighted average of 16 of its known neighbors. Keep in mind that pixels are not raster scanned but are encoded (and therefore also decoded) in levels. When decoding the pixels of level L, the MLP decoder has already decoded all the pixels of all the preceding levels, and it can use their values (gray scales) to predict values of pixels of L. Figure 7.140a shows the situation when the MLP encoder processes the last level. Half the pixels have already been encoded in previous levels, so they will be known to the decoder when the last level is decoded. The encoder can therefore use a diamond-shaped group of 4 Χ 4 pixels (shown in black) from previous levels to predict the pixel at the center of the group. This group becomes the context of the pixel. Compression methods that scan the image in raster order can use only pixels above and to the left of pixel P to predict P. Because of the progressive nature of MLP, it can use a symmetric context, which produces more accurate predictions. On the other hand, the pixels of the context are not near neighbors and may even be (in early levels) far from the predicted pixel. Table 7.141 shows the 16 weights used for a group. They are calculated by bicubic polynomial interpolation (Section 7.25.4) and are normalized such that they add up to 1. (Notice that in Table 7.141a the weights are not normalized; they add up to 256. When these integer weights are used, the weighted sum should be divided by 256.) To predict a pixel near an edge, where some of the 16 neighbors are missing (as in Figure 7.140b), only those neighbors that exist are used, and their weights are renormalized, to bring their sum to 1. 1 -9 -9 -9 81 -9 1 81 81 1 -9 81 -9 -9 -9 1 0.0039 -0.0351 -0.0351 -0.0351 0.3164 -0.0351 0.0039 0.3164 0.3164 0.0039 -0.0351 0.3164 -0.0351 -0.0351 -0.0351 0.0039 Table 7.141: 16 Weights. (a) Integers. (b) Normalized.  Exercise 7.36: Why do the weights have to add up to 1?  Exercise 7.37: Show how to renormalize the six weights needed to predict the pixel at the bottom left corner of Figure 7.140b. The encoder predicts all the pixels of a level by using the diamond-shaped group of 4Χ4 (or fewer) “older” pixels around each pixel of the level. This is the image model used by MLP. It’s hard to predict, especially the future. —Niels Bohr 622 7. Image Compression 7.25.3 Error Modeling Assume that the weighted sum of the 16 near-neighbors of pixel P equals R. Thus, R is the value predicted for P. The prediction error, E, is simply the difference R - P. Assuming an image with 16 gray levels (four bits per pixel) the largest value of E is 15 (when R = 15 and P = 0) and the smallest is -15. Depending on the image, we can expect most of the errors to be small integers, either zero or close to it. Few errors will be ±15 or close to that. Experiments with a large number of images (see, for example, [Netravali and Limb 80]) have produced the error distribution shown in Figure 7.143a. This is a symmetric, narrow curve, with a sharp peak, indicating that most errors are small and are therefore concentrated at the top. Such a curve has the shape of the wellknown Laplace distribution (Section 10.2.1) with mean 0. This statistical distribution is similar to the normal (Gaussian) distribution but is narrower. It fits the error curve better than the normal distribution because most error values are small. x V 0 2 4 6 8 10 3: 0.408248 0.0797489 0.015578 0.00304316 0.00059446 0.000116125 4: 0.353553 0.0859547 0.020897 0.00508042 0.00123513 0.000300282 5: 0.316228 0.0892598 0.025194 0.00711162 0.00200736 0.000566605 1,000: 0.022360 0.0204475 0.018698 0.0170982 0.0156353 0.0142976 Table 7.142: Some Values of the Laplace Distribution with V = 3, 4, 5, and 1,000. Table 7.142 shows some values for the Laplace distributions with m = 0 and V = 3, 4, 5, and 1,000. Figure 7.143b shows the graphs of the first three of those. It is clear that as V grows, the graph becomes lower and wider, with a less-pronounced peak. The factor 1/.2V is included in the definition of the Laplace distribution in order to scale the area under the curve of the distribution to 1. Because of this, it is easy to use the curve of the distribution to calculate the probability of any error value. Figure 7.143c shows a gray strip, 1 unit wide, under the curve of the distribution, centered at an error value of k. The area of this strip equals the probability of any error E having the value k. Mathematically, the area is the integral PV (k) = 6 k+.5 k-.5 1 .2V exp7-42 V |x|8 dx, (7.40) and this is the key to encoding the errors. With 4-bit pixels, error values are in the range [-15, +15]. When an error k is obtained, the MLP encoder encodes it arithmetically with a probability computed by Equation (7.40). In practice, both encoder and decoder should have a table with all the possible probabilities precomputed. The only remaining point to discuss is what value of the variance V should be used in Equation (7.40). Both encoder and decoder need to know this value. It is clear, from Figure 7.143b, that using a large variance (which corresponds to a low, flat distribution) results in too low a probability estimate for small error values k. The arithmetic encoder would produce an unnecessarily long code in such a case. On the 7.25 MLP 623 -20 -10 0 10 20 3 4 5 x .3 .4 L(v,x) x x x x x x x x x x x x x x x x x x x x x x xx x x x x x x x x x x x -15 -8 0 8 15 error number k (a) (b) (c) of errors Figure 7.143: (a) Error Distribution. (b) Laplace Distributions. (c) Probability of k. 624 7. Image Compression other hand, using a small variance (which corresponds to a high, narrow distribution) would allocate too low probabilities to large error values k. The choice of variance is therefore important. An ideal method to estimate the variance should assign the best variance value to each error and should involve no overhead (i.e., no extra data should be written on the compressed stream to help the decoder estimate the variances). Here are some approaches to variance selection: 1. The Laplace distribution was adopted as the MLP error distribution after many experiments with real images. The distribution obtained by all those images has a certain value of V , and this value should be used by MLP. This is a simple approach, which can be used in fast versions of MLP. However, it is not adaptive (since it always uses the same variance), and therefore does not result in best compression performance for all images. 2. (A two-pass compression job.) Each compression should start with a pass where all the error values are computed and their variance calculated (see below). This variance should be used, after the first pass, to compute a table of probabilities from Equation (7.40). The table should be used in the second pass where the error values are encoded and written on the compressed stream. Compression performance would be excellent, but any two-pass job is slow. Notice that the entire table can be written at the start of the compressed stream, thereby greatly simplifying the task of the decoder.  Exercise 7.38: Show an example where this approach is practical (i.e., when slow encoding is unimportant but a fast decoder and excellent compression are important). 3. Every time an error is obtained and is about to be arithmetically coded, use some method to estimate the variance associated with that error. Quantize the estimate and use a number of precomputed probability tables, one for each quantized variance value, to compute the probability of the error. This is computationally intensive but may be a good compromise between approaches 1 and 2 above. We now need to discuss how the variance of an error can be estimated, and we start with an explanation of the concept of variance. Variance is a statistical concept defined for a sequence of values a1, a2, . . . , an. It measures how elements ai vary by calculating the differences between them and the average A of the sequence, which is defined, as usual, as A = (1/n)ai. A small variance is the reason why the curves of Figure 7.143b that correspond to smaller variances are narrower; their values are concentrated near the average, which in this case is zero. The sequence (5, 5, 5), for example, has average 5 and variance 0, because every element of the sequence equals the average. The sequence (0, 5, 10) also has average 5 but should have a nonzero variance, because two of its elements differ from the average. In general, the variance of the sequence ai is defined as the nonnegative quantity V = .2 = E(ai - A)2 = 1 n n 1 (ai - A)2, so the variance of (0, 5, 10) is [(0-5)2 +(5-5)2+(10-5)2]/3 = 50/3. Statisticians also use a quantity called standard deviation (denoted by .) that is defined as the square root of the variance. 7.25 MLP 625 We now discuss several ways to estimate the variance of a prediction error E. 3.1. Equation (7.40) gives the probability of an error E with value k, but this probability depends on V . We can consider PV (k) a function of the two variables V and k, and find the optimal value of V by solving the equation .PV (k)/.V = 0. The solution is V = 2k2, but this method is not practical, because the decoder does not know k (it is trying to decode k so it can find the value of pixel P), and thus cannot mirror the operations of the encoder. It is possible to write the values of all the variances on the compressed stream, but this would significantly reduce the compression ratio. This method can be used to encode a particular image in order to find the best compression ratio of the image and compare it to what is achieved in practice. 3.2. While the pixels of a level are being encoded, consider their errors E to be a sequence of numbers, and find its variance V . Use V to encode the pixels of the next level. The number of levels is never very large, so all the variance values can be written (arithmetically encoded) on the compressed stream, resulting in fast decoding. The variance used to encode the first level should be a user-controlled parameter whose value is not critical, because that level contains just a few pixels. Since MLP quantizes a variance to one of 37 values (see below), which is why each variance written on the compressed stream is encoded in just log2 37 . 5.21 bits, a negligible overhead. The obvious disadvantage of this method is that it disregards local concentrations of identical or very similar pixels in the same level. 3.3. (Similar to 3.2.) While the pixels of a level are being encoded, collect the prediction errors of each block of bΧb pixels and use them to compute a variance that will be used to encode the pixels inside this block in the next level. The variance values for a level can also be written on the compressed stream following all the encoded errors for that level, so the decoder could use them without having to compute them. Parameter b should be adjusted by experiments, and the authors recommend the value b = 16. This method entails significant overhead and may therefore degrade compression performance. 3.4. (This is a later addition to MLP; see [Howard and Vitter 92b].) A variability index is computed, by both the encoder and decoder, for each pixel. This index should depend on the amount by which the pixel differs from its near neighbors. The variability indexes of all the pixels in a level are then used to adaptively estimate the variances for the pixels, based on the assumption that pixels with similar variability index should use Laplace distributions with similar variances. The method proceeds in the following steps: 1. Variability indexes are calculated for all the pixels of the current level, based on values of pixels in the preceding levels. This is done by the encoder and is later mirrored by the decoder. After several tries, the developers of MLP have settled on a simple way to calculate the variability index. It is calculated as the variance of the four nearest neighbors of the current pixel (the neighbors are from preceding levels, so the decoder can mirror this operation). 2. The variance estimate V is set to some initial value whose choice is not critical, as V is going to be updated often later. The decoder chooses this value in the same way. 3. The pixels of the current level are sorted in variability index order. The decoder can mirror this even though it still does not have the values of these pixels (the decoder has already calculated the values of the variability index in step 1, because they depend on pixels of previous levels). 626 7. Image Compression 4. The encoder loops over the sorted pixels in decreasing order (from large variability indexes to small ones). For each pixel: 4.1. The encoder calculates the error E of the pixel and sends E and V to the arithmetic encoder. The decoder mirrors this step. It knows V , so it can decode E. 4.2. The encoder updates V by V ‹ fΧV + (1 - f)E2, where f is a smoothing parameter (experience suggests a large value, such as 0.99, for f). This is how V is adapted from pixel to pixel, using the errors E. Because of the large value of f, V is decreased in small steps. This means that latter pixels (those with small variability indexes) will get small variances assigned. The idea is that compressing pixels with large variability indexes is less sensitive to accurate values of V . As the loop progresses, V gets assigned more accurate values, and these are used to compress pixels with small variability indexes, which are more sensitive to variance values. Notice that the decoder can mirror this step, since it has already decoded E in step 4.1. Notice also that the arithmetic encoder writes the encoded error values on the compressed stream in decreasing variability index order, not row by row. The decoder can mirror this too, since it has already sorted the pixels in this order in step 3. This method gives excellent results but is even more computationally intensive than the original MLP (end of method 3.4). Using one of the four methods above, variance V is estimated. Before using V to encode error E, V is quantized to one of 37 values as shown in Table 7.145. For example, if the estimated variance value is 0.31, it is quantized to 7. The quantized value is then used to select one of 37 precomputed probability tables (in our example Table 7, precomputed for variance value 0.290, is selected) prepared using Equation (7.40), and the value of error E is used to index that table. (The probability tables are not shown here.) The value retrieved from the table is the probability that’s sent to the arithmetic encoder, together with the error value, to arithmetically encode error E. MLP is therefore one of the many compression methods that implement a model to estimate probabilities and use arithmetic coding to do the actual compression. Table 7.144 is a pseudo-code summary of MLP encoding. 7.25.4 Interpolating Polynomials This section shows how to predict the value of a pixel from those of 16 of its near neighbors by means of a two-dimensional interpolating polynomial. The results are used in Table 7.141. We start with an intuitive discussion of the term interpolation. Given two numbers a and b, their average (a + b)/2 is always located midway between them, so we can use the average to interpolate them. However, given four numbers a, b, c, and d, their average (a+b+c+d)/4 is not a good interpolation, because it is not located “midway” between the four. A simple example is the four numbers 1, 1, 1, and 100. Their average is close to 25, so it is nowhere “in the middle” of the four numbers. Interpolating four numbers is therefore done by (1) converting the numbers to two-dimensional points, (2) calculating a smooth curve that passes through the points, and (3) finding the midpoint of the curve. 7.25 MLP 627 for each level L do for every pixel P in level L do Compute a prediction R for P using a group from level L-1; Compute E=R-P; Estimate the variance V to be used in encoding E; Quantize V and use it as an index to select a Laplace table LV; Use E as an index to table LV and retrieve LV[E]; Use LV[E] as the probability to arithmetically encode E; endfor; Determine the pixels of the next level (rotate & scale); endfor; Table 7.144: MLP Encoding. Variance Var. Variance Var. Variance Var. range used range used range used 0.005–0.023 0.016 2.882–4.053 3.422 165.814–232.441 195.569 0.023–0.043 0.033 4.053–5.693 4.809 232.441–326.578 273.929 0.043–0.070 0.056 5.693–7.973 6.747 326.578–459.143 384.722 0.070–0.108 0.088 7.973–11.170 9.443 459.143–645.989 540.225 0.108–0.162 0.133 11.170–15.627 13.219 645.989–910.442 759.147 0.162–0.239 0.198 15.627–21.874 18.488 910.442–1285.348 1068.752 0.239–0.348 0.290 21.874–30.635 25.875 1285.348–1816.634 1506.524 0.348–0.502 0.419 30.635–42.911 36.235 1816.634–2574.021 2125.419 0.502–0.718 0.602 42.911–60.123 50.715 2574.021–3663.589 3007.133 0.718–1.023 0.859 60.123–84.237 71.021 3663.589–5224.801 4267.734 1.023–1.450 1.221 84.237–118.157 99.506 5224.801–7247.452 6070.918 1.450–2.046 1.726 118.157–165.814 139.489 7247.452–10195.990 8550.934 2.046–2.882 2.433 Table 7.145: Thirty-Seven Quantized Variance Values. Any numbers a, b, c, and d can be converted to the points (1, a), (2, b), (3, c), and (4, d). It is intuitively clear that the midpoint (x, y) of a smooth curve that passes through those points is a good candidate for the title “the interpolation of the four points.” The y coordinate becomes the interpolation of the four numbers, and the x coordinate is ignored. This method is called one-dimensional interpolation. It can be extended to more than four numbers, and also to pixels, where it becomes two-dimensional interpolation. As mentioned before, we want to use a group of 16 neighboring pixels to predict the value of a pixel at the center of the group. The main idea is to consider the 16 neighbor pixels a set of 4 Χ 4 equally-spaced points on a surface (where the value of a pixel is interpreted as the height of the surface) and to derive a polynomial function P(u,w) that passes through all 16 points. Graphically, P(u,w) can be thought of as a surface. 628 7. Image Compression The value of the pixel at the center of the 4Χ4 group can then be predicted by calculating the height of the center point P(0.5, 0.5) of the surface. Mathematically, this surface is the two-dimensional polynomial interpolation of the 16 points. 7.25.5 One-Dimensional Interpolation A surface can be viewed as an extension of a curve, so we start by deriving a onedimensional polynomial (a curve) that interpolates four points, then extend it to a two-dimensional polynomial (a surface) that interpolates a grid of 4 Χ 4 points. Given four points P1, P2, P3, and P4 we look for a polynomial that will pass through them. In general, a polynomial of degree n in x is defined (Section 6.32) as the function Pn(x) = n i=0 aixi = a0 + a1x + a2x2 + · · · + anxn, (7.41) where ai are the n+1 coefficients of the polynomial and the parameter x is a real number. The one-dimensional interpolating polynomial that is of interest to us is special, and differs from the definition above in two respects 1. This polynomial goes from point P1 to point P4. Its length is finite, and it is therefore better to describe it as the function Pn(t) = n i=0 aiti = a0 + a1t + a2t2 + · · · + antn; where 0 . t . 1. This is the parametric representation of a polynomial. We want this polynomial to go from P1 to P4 when the parameter t is varied from 0 to 1. 2. The only given data are the four points and we have to use them to calculate all n + 1 coefficients of the polynomial. This suggests the value n = 3 (a polynomial of degree 3, a cubic polynomial; one that has four coefficients). The idea is to set up and solve four equations, with the four coefficients as the unknowns, and with the four points as known quantities. Thus, we use the notation (T indicates transpose) P(t) = at3 + bt2 + ct + d = (t3, t2, t, 1)(a, b, c, d)T = T(t) · A. (7.42) The four coefficients a, b, c, d are shown in boldface because they are not numbers. Keep in mind that the polynomial has to pass through the given points, so the value of P(t) for any t must be the three coordinates of a point. Each coefficient should therefore be a triplet. T(t) is the row vector (t3, t2, t, 1), and A is the column vector (a, b, c, d)T . Calculating the curve therefore involves finding the values of the four unknowns a, b, c, and d. P(t) is called a parametric cubic (or PC) polynomial. It turns out that degree 3 is the smallest one that is still useful for an interpolating polynomial. A polynomial of degree 1 has the form P1(t) = ct + d and is therefore a straight line, so it can be used only in special cases. A polynomial of degree 2 (quadratic) has the form P2(t) = bt2+ct+d and is a conic section, so it can take only a few different shapes. A polynomial of degree 3 (cubic) is therefore the simplest one that can take on complex shapes, and can also be a space curve. 7.25 MLP 629  Exercise 7.39: Prove that a quadratic polynomial must be a plane curve. Our ultimate problem is to interpolate pixels. Pixels are always spaced uniformly, so we assume that the two interior points P2 and P3 are equally spaced between P1 and P4. The first point P1 is the start point P(0) of the polynomial, the last point, P4 is the endpoint P(1), and the two interior points P2 and P3 are the two equally-spaced interior points P(1/3) and P(2/3) of the polynomial. We therefore write P(0) = P1, P(1/3) = P2, P(2/3) = P3, P(1) = P4, or a(0)3 + b(0)2 + c(0) + d = P1, a(1/3)3 + b(1/3)2 + c(1/3) + d = P2, a(2/3)3 + b(2/3)2 + c(2/3) + d = P3, a(1)3 + b(1)2 + c(1) + d = P4. These equations are easy to solve, and the solutions are: a = -9/2P1 + 27/2P2 - 27/2P3 + 9/2P4, b = 9P1 - 45/2P2 + 18P3 - 9/2P4, c = -11/2P1 + 9P2 - 9/2P3 + P4, d = P1. Substituting into Equation (7.42) gives P(t) =(-9/2P1 + 27/2P2 - 27/2P3 + 9/2P4)t3 + (9P1 - 45/2P2 + 18P3 - 9/2P4)t2 + (-11/2P1 + 9P2 - 9/2P3 + P4)t + P1, which, after rearranging, becomes P(t) =(-4.5t3 + 9t2 - 5.5t + 1)P1 + (13.5t3 - 22.5t2 + 9t)P2 + (-13.5t3 + 18t2 - 4.5t)P3 + (4.5t3 - 4.5t2 + t)P4 =G1(t)P1 + G2(t)P2 + G3(t)P3 + G4(t)P4 =G(t) · P, (7.43) where G1(t) = (-4.5t3 + 9t2 - 5.5t + 1), G3(t) = (-13.5t3 + 18t2 - 4.5t), G2(t) = (13.5t3 - 22.5t2 + 9t), G4(t) = (4.5t3 - 4.5t2 + t); (7.44) P is the column (P1,P2,P3,P4)T , and G(t) is the row vector G1(t),G2(t),G3(t),G4(t) . The functions Gi(t) are called blending functions, because they create any point on the curve as a blend of the four given points. Note that they add up to 1 for any value of 630 7. Image Compression t. This property must be satisfied by any set of blending functions, and such functions are called barycentric. We can also write G1(t) = (t3, t2, t, 1)(-4.5, 9,-5.5, 1)T and, similarly, for G2(t), G3(t), and G4(t). In matrix notation this becomes G(t) = (t3, t2, t, 1)... -4.5 13.5 -13.5 4.5 9.0 -22.5 18 -4.5 -5.5 9.0 -4.5 1.0 1.0 0 0 0 ... = T(t) ·N. (7.45) The curve can now be written P(t) = G(t) · P = T(t)·N·P. N is called the basis matrix, and P is the geometry vector. From Equation (7.42) we know that P(t) = T(t) · A, so we can write A = N· P. Equation (8.17) illustrates an application of this interpolating polynomial for image compression. The word barycentric is derived from barycenter, meaning “center of gravity,” because such weights are used to calculate the center of gravity of an object. Barycentric weights have many uses in geometry in general, and in curve and surface design in particular. Given the four points, the interpolating polynomial can be calculated in two steps: 1. Set-up the equation A = N· P and solve it for A = (a, b, c, d)T . 2. The polynomial is P(t) = T(t) · A. 7.25.6 Example (This example is in two dimensions, each of the four points Pi and each of the four coefficients a, b, c, and d is a pair. For three-dimensional curves, the method is the same, except that triplets should be used, instead of pairs.) Given the four two-dimensional points P1 = (0, 0), P2 = (1, 0), P3 = (1, 1), and P4 = (0, 1), we set up the equation ... abcd ... = A = N· P =... -4.5 13.5 -13.5 4.5 9.0 -22.5 18 -4.5 -5.5 9.0 -4.5 1.0 1.0 0 0 0 ... ... (0, 0) (1, 0) (1, 1) (0, 1) ... , which is easy to solve a = -4.5(0, 0) + 13.5(1, 0) - 13.5(1, 1) + 4.5(0, 1) = (0,-9), b = 19(0, 0) - 22.5(1, 0) + 18(1, 1) - 4.5(0, 1) = (-4.5, 13.5), c = -5.5(0, 0) + 9(1, 0) - 4.5(1, 1) + 1(0, 1) = (4.5,-3.5), d = 1(0, 0) - 0(1, 0) + 0(1, 1) - 0(0, 1) = (0, 0). 7.25 MLP 631 Thus P(t) = T · A = (0,-9)t3 + (-4.5, 13.5)t2 + (4.5,-3.5)t. It is now easy to calculate and verify that P(0) = (0, 0) = P1, and P(1/3) = (0,-9)1/27 + (-4.5, 13.5)1/9 + (4.5,-3.5)1/3 = (1, 0) = P2, P(1) = (0,-9)13 + (-4.5, 13.5)12 + (4.5,-3.5)1 = (0, 1) = P4.  Exercise 7.40: Calculate P(2/3) and verify that it is equal to P3.  Exercise 7.41: Imagine the circular arc of radius one in the first quadrant (a quarter circle). Write the coordinates of the four points that are equally spaced on this arc. Use the coordinates to calculate a PC interpolating polynomial approximating this arc. Calculate point P(1/2). How far does it deviate from the midpoint of the true quarter circle? The main advantage of this method is its simplicity. Given the four points, it is easy to calculate the PC polynomial that passes through them.  Exercise 7.42: This method makes sense if the four points are (at least approximately) equally spaced along the curve. If they are not equally spaced, the following may be done: Instead of using 1/3 and 2/3 as the intermediate values, the user may specify values ., ί such that P2 = P(.) and P3 = P(ί). Generalize Equation (7.45) such that it depends on . and ί. 7.25.7 Two-Dimensional Interpolation The PC polynomial, Equation (7.42), can easily be extended to two dimensions by means of a technique called Cartesian product. The polynomial is generalized from a cubic curve to a bicubic surface. A one-dimensional PC polynomial has the form P(t) = 3 i=0 aiti. Two such curves, P(u) and P(w), can be combined by means of this technique to form the surface: P(u,w) = 3 i=0 3 j=0 aijuiwj = a33u3w3 + a32u3w2 + a31u3w + a30u3 + a23u2w3 + a22u2w2 + a21u2w + a20u2 + a13uw3 + a12uw2 + a11uw + a10u + a03w3 + a02w2 + a01w + a00 = (u3, u2, u, 1)... a33 a32 a31 a30 a23 a22 a21 a20 a13 a12 a11 a10 a03 a02 a01 a00 ... ... w3 w2 w1 ... , where 0 . u,w . 1. (7.46) This is a double cubic polynomial (hence the name bicubic) with 16 terms, where each of the 16 coefficients aij is a triplet. Note that the surface depends on all 16 coefficients. Any change in any of them produces a different surface. Equation (7.46) is the algebraic representation of a bicubic surface. In order to use it in practice, the 632 7. Image Compression 16 unknown coefficients have to be expressed in terms of the 16 known, equally-spaced points. We denote these points P03 P13 P23 P33 P02 P12 P22 P32 P01 P11 P21 P31 P00 P10 P20 P30. To determine the 16 unknown coefficients, we write 16 equations, each based on one of the given points: P(0, 0) = P00 P(0, 1/3) = P01 P(0, 2/3) = P02 P(0, 1) = P03 P(1/3, 0) = P10 P(1/3, 1/3) = P11 P(1/3, 2/3) = P12 P(1/3, 1) = P13 P(2/3, 0) = P20 P(2/3, 1/3) = P21 P(2/3, 2/3) = P22 P(2/3, 1) = P23 P(1, 0) = P30 P(1, 1/3) = P31 P(1, 2/3) = P32 P(1, 1) = P33. Solving, substituting the solutions in Equation (7.46), and simplifying produces the geometric representation of the bicubic surface P(u,w) = (u3, u2, u, 1)N... P33 P32 P31 P30 P23 P22 P21 P20 P13 P12 P11 P10 P03 P02 P01 P00 ... NT... w3 w2 w1 ... , (7.47) where N is the Hermite matrix of Equation (7.45). The surface of Equation (7.47) can now be used to predict the value of a pixel as a polynomial interpolation of 16 of its near neighbors. All that is necessary is to substitute u = 0.5 and w = 0.5. The following Mathematica code Clear[Nh,P,U,W]; Nh={{-4.5,13.5,-13.5,4.5},{9,-22.5,18,-4.5}, {-5.5,9,-4.5,1},{1,0,0,0}}; P={{p33,p32,p31,p30},{p23,p22,p21,p20}, {p13,p12,p11,p10},{p03,p02,p01,p00}}; U={u^3,u^2,u,1}; W={w^3,w^2,w,1}; u:=0.5; w:=0.5; Expand[U.Nh.P.Transpose[Nh].Transpose[W]] does that and produces P(0.5, 0.5) = 0.00390625P00 - 0.0351563P01 - 0.0351563P02 + 0.00390625P03 - 0.0351563P10 + 0.316406P11 + 0.316406P12 - 0.0351563P13 - 0.0351563P20 + 0.316406P21 + 0.316406P22 - 0.0351563P23 + 0.00390625P30 - 0.0351563P31 - 0.0351563P32 + 0.00390625P33, where the 16 coefficients are the ones used in Table 7.141. 7.26 Adaptive Golomb 633  Exercise 7.43: How can this method be used in cases where not all 16 points are known?  Exercise 7.44: The center point of the surface is calculated as a weighted sum of the 16 equally-spaced data points. It makes sense to assign small weights to points located away from the center, but our result assigns negative weights to eight of the 16 points. Explain the meaning of negative weights and show what role they play in interpolating the center of the surface. Readers who find it difficult to follow the details above should compare the way two-dimensional polynomial interpolation is presented here to the way it is discussed by [Press et al. 88]. The following quotation is from page 125: “. . . The formulas that obtain the c’s from the function and derivative values are just a complicated linear transformation, with coefficients which, having been determined once, in the mists of numerical history, can be tabulated and forgotten.” 7.26 Adaptive Golomb The principle of run-length encoding is simple. Given a bi-level image, it is scanned row by row and alternating run lengths of black and white pixels are computed. Each run length is in the interval [1, r], where r is the row size. In principle, the run lengths constitute the compressed image, but in practice we need to write them on the compressed stream such that the decoder will be able to read them unambiguously. This is done by replacing each run length by a prefix code and writing the codes on the compressed stream without any separators. This section shows that the Golomb codes (Section 3.24) are ideal for this application. The Golomb codes depend on a parameter m, so we also propose a simple adaptive algorithm for estimating m for each run length in order to get the best compression in a one-pass algorithm. Figure 7.146 is a simple finite-state machine (see Section 11.8 for more information) that summarizes the process of reading black and white pixels from an image. At any time, the machine can be in one of two states: a 0 (white) or a 1 (black). If the current state is 0 (i.e., the last pixel input was white), then the probability that the next state will be a 1 (the next pixel input will be black) is denoted by ., which implies that the probability that the next state will also be 0 is 1 - .. Suppose we are at state 0. The probability pw(1) of a sequence of exactly one white pixel is the probability . of a transition to state 1. Similarly, the probability pw(2) of a sequence of exactly two white pixels is the probability of returning once to state 0, then switching to state 1. Since these events are independent, the probability is the product (1 - .).. In general, the probability pw(n) of a run length of n white pixels is (1 - .)n-1. (this can be proved by induction). Thus, we conclude that pw(n) and, by symmetry, also pb(n) = (1 - ί)n-1ί are distributed geometrically. This analysis is known as the Capon Markov-1 model of the run lengths of a bi-level image [Capon 59] and it justifies (as explained in Section 3.24) the use of Golomb codes to compress the run lengths. Once this is understood, it is easy to derive a simple, adaptive process that estimates the best Golomb parameter m for each run length, computes the Golomb code for the 634 7. Image Compression 0 1 . ί 1-. 1-ί Figure 7.146: A Finite-State Machine for Reading Bi-level Pixels. run length, and updates a table that’s used to obtain better estimates for the remainder of the input. Each run length is an integer in the interval [1, r], where r is the row size. The first step is to decrement the run length by 1, since the Golomb codes are for the nonnegative (as opposed to the positive) integers. Thus, we denote the modified run length of l pixels of color c by (c, l), where c is either b or w and l is in the interval [0, r - 1]. A table S[b:w, 0:r - 1] of two rows indexed by b and w and r columns indexed from 0 to r - 1 is maintained with counts of past run lengths and is used to estimate the probability p of each run length. Assume that the current run length is (c, l), then p is computed as entry S[c, l] divided by the sum of all entries of S. p is then used to compute the best Golomb parameter m as the integer nearest -1/ log2(1 - p). Run length (c, l) is compressed by computing the Golomb code for the integer l with the parameter m, and table S is updated by incrementing element S[c, l] by 1. This process can be reversed by the decoder while maintaining the same table. There is the question of how to initialize S. In principle, all of its entries should be initialized to zero, but this would produce probabilities of zero, which cannot be used to compute m values. This is the zero-probability problem, introduced in Section 5.14. A reasonable solution is to initialize each table entry to a default value such as 1 and allow the user to change this value according to the nature of the image. This makes sense because even though bi-level is the simplest image type, there is a wide variety of such images. A page of printed text, for example, normally results, once digitized, in 5–6% of black pixels and with very short runs of those pixels (this figure is used by printer manufacturers to estimate the life of print cartridges). The values of . and ί for such a page are about 0.02–0.03 and 0.34–0.35, respectively. In the other extreme, we find traditional woodcuts. When a woodcut is printed and digitized in black and white, it often results in a large majority (and with long runs) of black pixels. A woodcut is art carved on (and printed directly from) a flat piece of wood. The parts to be printed remain flat with the surface of the wood and the remaining parts should be removed by the artist. A roller is soaked with ink and is rolled on the wood so it covers the surface, but not the carved parts, with ink. A sheet of paper is then placed on the wood and is pressed, either by hand or by a roller, to transfer the ink from the wood. The resulting print on the paper is a mirror image of that on the wood, which makes it especially difficult to include text in a woodcut. When printed on paper, a woodcut tends to have large black areas because any white (nonprinting) areas have to be removed from the wood, but black areas are those parts of the surface of the wood that are untouched by the artist. It is possible to have color woodcuts. The artist starts with several flat pieces of 7.27 PPPM 635 wood, each for one color. In the piece for red, for example, the red parts of the image are untouched and the other parts are removed. The same sheet of paper is then pressed to each piece of wood in turn and is left to dry between presses. Because modern life is fast, many woodcut artists remove the nonprinting areas from the wood by sandblasting instead of slow hand carving (after covering the printing areas with a metal or plastic shield). The art of carving a woodcut is called xylography. 7.27 PPPM The reader should review the PPM method, Section 5.14, before reading this section. The PPPM method uses the ideas of MLP (Section 7.25). It is also (remotely) related to the context-based image compression method of Section 7.22. PPM encodes a symbol by comparing its present context to other similar contexts and selecting the longest match. The context selected is then used to estimate the symbol’s probability in the present context. This way of context matching works well for text, where we can expect strings of symbols to repeat exactly, but it does not work well for images, because a digital image is often the result of digitizing an analog image. Assume that the current pixel has intensity 118 and its context is the two neighboring pixels with values 118 and 120. It is possible that 118 was never seen in the past with the context 118, 120 but was seen with contexts 119, 120 and 118, 121. Clearly, these contexts are close enough to the current one to justify using one of them. Once a closely matching context has been found, it is used to estimate the variance (not the probability) of the current prediction error. This idea serves as one principle of the Prediction by Partial Precision Matching (PPPM) method [Howard and Vitter 92a]. The other principle is to use the Laplace distribution to estimate the probability of the prediction error, as done in MLP. Figure 7.147 shows how prediction is done in PPPM. Pixels are raster-scanned, row by row. The two pixels labeled C are used to predict the one labeled P. The prediction R is simply the rounded average of the two C pixels. Pixels in the top or left edges 636 7. Image Compression . . . . . . S C S . . C P Figure 7.147: Prediction and Variance-Estimation Contexts for PPPM. are predicted by one neighbor only. The top-left pixel of the image is encoded without prediction. After predicting the value of P, the encoder calculates the error E = R - P and uses the Laplace distribution to estimate the probability of the error, as in MLP. The only remaining point to discuss is how PPPM estimates the variance of the particular Laplace distribution that should be used to obtain the probability of E. PPPM uses the four neighbors labeled C and S in Figure 7.147. These pixels have already been encoded, so their values are known. They are used as the variance-estimation context of P. Assume that the four values are 3, 0, 7, and 5, expressed as 4-bit numbers. They are combined to form the 16-bit key 0011|0000|0111|0101, and the encoder uses a hash table to find all the previous occurrences of this context. If this context occurred enough times in the past (more than the value of a threshold parameter), the statistics of these occurrences are used to obtain a mean m and a variance V. If the context did not occur enough times in the past, the least-significant bit of each of the four values is dropped to obtain the 12-bit key 001|000|011|010, and the encoder hashes this value. Thus, the encoder iterates in a loop until it finds m and V. (It turns out that using the errors of the C and S pixels as a key, instead of their values, produces slightly better compression, so this is what PPPM actually does.) Once m and V have been obtained, the encoder quantizes V and uses it to select one of 37 Laplace probability tables, as in MLP. The encoder then adds E+m and sends this value to be arithmetically encoded with the probability obtained from the Laplace table. To update the statistics, the PPPM encoder uses a lazy approach. It updates the statistics of the context that is actually used and also updates, if applicable, the context with one additional bit of precision. One critical point is the number of times a context had to be seen in the past to be considered meaningful and not random. The PPMB method, Section 5.14.4, “trusts” a context if it has been seen twice. For an image, a threshold of 10–15 is more reasonable. 7.28 CALIC Sections 7.22 through 7.27 describe context-based image compression methods that have one feature in common: They determine the context of a pixel based on some of its neighbor pixels that have already been seen and processed. Normally, these are some of the pixels above and to the left of the current pixel, which leads to asymmetric context. It seems intuitive that a symmetric context, one that predicts the current pixel from pixels all around it, would produce better compression, so attempts have been made to develop image compression methods that employ such contexts. The MLP method, Section 7.25, provides an interesting twist to the problem of symmetric context. The CALIC method of this section takes a different approach. The 7.28 CALIC 637 name CALIC ([Wu 95] and [Wu 96]) stands for Context-based, Adaptive, Lossless Image Compression. It performs three passes to create a symmetric context around the current pixel, and it uses quantization to reduce the number of possible contexts to something manageable. The method has been developed for compressing grayscale images (where each pixel is a c-bit number representing a shade of gray), but like any other method for grayscale images, it can handle a color image by separating it into three color components and treating each component as a grayscale image. 7.28.1 Three Passes We start with an image I[i, j] that consists of H rows and W columns of pixels. Both encoder and decoder perform three passes over the image. The first pass calculates averages of pairs of pixels. It looks only at pixels I[i, j] where i and j have the same parity (i.e., both are even or both are odd). The second pass uses these averages to actually encode the same pixels. The third pass uses the same averages plus the pixels of the second pass to encode all pixels I[i, j] where i and j have different parities (one is odd and the other is even). The first pass calculates the W/2 Χ H/2 values ΅[i, j] defined by ΅[i, j] = (I[2i, 2j] + I[2i + 1, 2j + 1])/2, for 0 . i < H/2, 0 . j min , min , max ) ; encode( knot->max , min , max ) ; if ( knot->min == knot->max ) return ; min = knot->min ; max= knot->max ; if ( knot->depth < Log2(N) ) { Compact_quadtree(knot->square[ 0 ],min,max) ; Compact_quadtree(knot->square[ 1 ],min,max) ; Compact_quadtree(knot->square[ 2 ],min,max) ; Compact_quadtree(knot->square[ 3 ],min,max) ; } else { // knot->depth == Log2(N) e pixel level int slc = 0 ; encode( (knot->square[ 0 ])->pix , min , max ); if ((knot->square[ 0 ])->pix == min) slc=1; else if ((knot->square[ 0 ])->pix==max) slc=2; encode( (knot->square[ 1 ])->pix , min , max ); if ((knot->square[ 1 ])->pix == min) slc |= 1; else if ((knot->square[ 1 ])->pix==max) slc |= 2; switch( slc ) { case 0: encode(((knot->square[2])->pix==max),0,1); return ; case 1: if ((knot->square[2])->pix==max) { 7.34 Quadtrees 673 encode(1,0,1); encode((knot->square[ 3 ])->pix,min,max); } else { encode(0,0,1); encode((knot->square[ 2 ])->pix,min,max); } return ; case 2: if ((knot->square[2])->pix==min) { encode(1,0,1); encode((knot->square[ 3 ])->pix,min,max); } else { encode(0,0,1); encode((knot->square[ 2 ])->pix,min,max); } return ; case 3: encode((knot->square[ 2 ])->pix,min,max); encode((knot->square[ 3 ])->pix,min,max); } } } An improvement The following improvement to the method of this section was communicated to me by its originator, Stephan Wolf, who also wrote these paragraphs. I was surprised by the fact that arithmetic encoding is suggested for the encode() function. Arithmetic coding is apparently the best method if individual symbols have different probabilities. But in this case, the paper seems to assume equal probabilities for all the individual pixel values in a particular encoded range. Thus, I had the idea for a very simple algorithm that uses exactly log2 n bits to encode a value for a given zero-relative range of values. I thought I would wait until I receive your book and see if it describes this algorithm, i.e., if this is an already-known technique. But I could not find any hints in your book or anywhere else (for example, the Internet) about the algorithm I devised. So may I present you with this algorithm for revisal. May I also add that the following ideas are all mine: Given a (zero-relative) range r of equally-distributed values v, i.e., all values have equal probabilities. In this paper, I denote a range/value pair as a tuple (r, v). Both encoder and decoder need to already know the range r in each step of encoding/decoding an individual value v. As an example, I use the following range/value pairs (9, 2), (199, 73), (123, 89), and (50, 43). The theoretical number of bits required to encode each individual value is log2(r+1) which translates (in rounded values using three decimal digits) to log2(9+1) = 3.32, log2(199 + 1) = 7.64, log2(123 + 1) = 6.95, and log2(50 + 1) = 5.67. Thus, a total 674 7. Image Compression maximum of 23.6 bits is required to encode all values in the example. A perfect encoder would therefore use at most 24 bits. I propose the algorithm e[n+1]=e[n]*(r[n]+1)+v[n] to encode the values in the range/value pairs above. The results are e[1] = 0Χ (9 + 1) + 2 = 2, e[2] = 2Χ (199 + 1) + 73 = 473, e[3] = 473 Χ (123 + 1) + 89 = 58,741, e[4] = 58,741 Χ (50 + 1) + 43 = 2,995,834. This value can be expressed in 22 bits as 1011011011011001111010. Decoding is quite similar (note that % denotes the modulo operation, i.e., the integral remainder of division) v[n+1] = d[n] % r d[n+1] = d[n] / r v[1] = 2,995,834 % (50+1) = 43 d[1] = 2,995,834 / (50+1) = 58,741 v[2] = 58,741 % (123+1) = 89 d[2] = 58,741 / (123+1) = 473 v[3] = 473 % (199+1) = 73 d[3] = 473 / (199+1) = 2 v[4] = 2 % (9+1) = 2 d[4] = 2 / (9+1) = 0 Note that any implementation of this encoding/decoding algorithm requires long integer arithmetic, similar to what is normally available in most cryptography libraries. 7.34.5 Prefix Compression Prefix compression is a variant of quadtrees proposed in [Anedda and Felician 88] (see also Section 11.5.5). We start with a 2nΧ2n image. Each quadrant in the quadtree of this image is numbered 0, 1, 2, or 3, a two-bit number. Each subquadrant has a two-digit (i.e., a four-bit) number, and each subsubquadrant receives a 3-digit number. As the quadrants get smaller, their numbers get longer. When this numbering scheme is carried down to individual pixels, the number of a pixel turns out to be n digits, or 2n bits, long. Prefix compression is designed for bi-level images containing text or diagrams where the number of black pixels is relatively small. It is not suitable for grayscale, color, or any image that contains many black pixels, such as a painting. The method is best explained by an example. Figure 7.172 shows the pixel numbering in an 8Χ8 image (i.e., n = 3) and also a simple 8Χ8 image consisting of 18 black pixels. Each pixel number is three digits long, and they range from 000 to 333. The first step is to use quadtree methods to figure the three-digit id numbers of the 18 black pixels. They are 000 101 003 103 030 121 033 122 123 132 210 301 203 303 220 221 222 223. The next step is to select a prefix value P. We select P = 2, a choice that’s justified below. The code of a pixel is now divided into P prefix digits followed by 3 - P suffix 7.34 Quadtrees 675 000 001 010 011 100 101 110 111 002 003 012 013 102 103 112 113 020 021 030 031 120 121 130 131 022 023 032 033 122 123 132 133 200 201 210 211 300 301 310 311 202 203 212 213 302 303 312 313 220 221 230 231 320 321 330 331 222 223 232 233 322 323 332 333 Figure 7.172: Example of Prefix Compression. digits. The last step goes over the sequence of black pixels and selects all the pixels with the same prefix. The first prefix is 00, so all the pixels that start with 00 are selected. They are 000 and 003. They are removed from the original sequence and are compressed by writing the token 00|1|0|3 on the output stream. The first part of this token is a prefix (00), the second part is a count (1), and the rest are the suffixes of the two pixels having prefix 00. Notice that a count of one implies two pixels. The count is always one less than the number of pixels being counted. Sixteen pixels now remain in the original sequence, and the first of them has prefix 10. The two pixels with this prefix are removed and compressed by writing the token 10|1|1|3 on the output stream. This continues until the original sequence is empty. The final result is the 9-token string 00|1|0|3, 10|1|1|3, 03|1|0|3, 12|2|1|2|3, 13|0|2, 21|0|0, 30|1|1|3, 20|0|3, 22|3|0|1|2|3 (without the commas). Such a string can be decoded uniquely, since each segment starts with a two-digit prefix, followed by a one-digit count c, followed by c+1 one-digit suffixes. In general, the prefix is P digits long, and the count and each suffix are n-P digits each. The maximum number of suffixes in a segment is therefore 4n-P . The maximum size of a segment is thus P +(n-P)+4n-P (n-P) digits. Each segment corresponds to a different prefix. A prefix has P digits, each between 0 and 3, so the maximum number of segments is 4P . The entire compressed string therefore occupies at most 4P P + (n - P) + 4n-P (n - P) = n · 4P + 4n(n - P) digits. To find the optimum value of P we differentiate the expression above with respect to P, d dP n · 4P + 4n(n - P) = n · 4P ln 4 - 4n, and set the derivative to zero. The solution is 4P = 4n n · ln 4, or P = log4  4n n · ln 4= 1 2 log2  4n n · ln 4. For n = 3 this yields P . 1 2 log2  43 3 Χ 1.386= log2 15.388 2 = 3.944/2 = 1.97. 676 7. Image Compression This is why P = 2 was selected earlier. A downside of this method is that some pixels may be assigned numbers with different prefixes even though they are near neighbors. This happens when they are located in different quadrants. An example is pixels 123 and 301 of Figure 7.172. Improvement: The count field was arbitrarily set to one digit (two bits). The maximum count is therefore 3 (= 112), i.e., four pixels. It is possible to have a variablesize count field containing a variable-length code, such as the unary code of Section 3.1. This way a single token could compress any number of pixels. 7.35 Quadrisection A quadtree exploits the redundancy in an image by examining smaller and smaller quadrants, looking for uniform areas. The method of quadrisection [Kieffer et al. 96a,b] is related to quadtrees, because it uses the quadtree principle for dividing an image into subimages. However, quadrisection does not divide an image into four parts, but rather reshapes the image in steps by increasing the number of its rows and decreasing the number of its columns. The method is lossless. It performs well for bi-level images and is illustrated here for such an image, but can also be applied to grayscale (and therefore to color) images. The method assumes that the original image is a 2kΧ2k square matrix M0, and it reshapes M0 into a sequence of matrices M1, M2, . . . , Mk+1 with fewer and fewer columns. These matrices naturally have more and more rows, and the quadrisection method achieves compression by searching for and removing duplicate rows. The more rows and the fewer columns a matrix has, the better the chance of having duplicate rows. The end result is a matrix Mk+1 with one column and not more than 64 rows. This matrix is treated as a short string and is written at the end of the compressed stream, to help in decoding and recovering the original image M0. The compressed stream must also contain information on how to reconstruct each matrix Mj-1 from its successor Mj . This information is in the form of an indicator vector denoted by Ij-1. Thus, the output consists of all the Ij vectors (each arithmetically encoded into a string wj ), followed by the bits of MT k+1 (the transpose of the last, single-column, matrix, in raw format). This string is preceded by a prefix that indicates the value of k, the sizes of all the wj ’s, and the size of MT k+1. Here is the encoding process in a little more detail. We assume that the original image is a 2kΧ2k square matrix M0 of 4k pixels, each a single bit. The encoder uses an operation called projection (discussed below) to construct a sequence of matrices M1, M2, up to Mk+1, such that Mj has 4k-j+1 columns. This implies that M1 has 4k-1+1 = 4k columns and therefore just one row. Matrix M2 has 4k-1 columns and therefore four rows. Matrix M3 has 4k-2 columns, but may have fewer than eight rows, since any duplicate rows are removed from M2 before M3 is constructed. The number of rows of M3 is four times the number of distinct rows of M2. Indicator vector I2 is constructed at the same time as M3 to indicate those rows of M2 that are duplicates. The size of I2 equals the number of rows of M2, and the elements of I2 are nonnegative integers. The decoder uses I2 to reconstruct M2 from M3. 7.35 Quadrisection 677 All the Ij indicator vectors are needed by the decoder. Since the size of Ij equals the number of rows of matrix Mj , and since these matrices have more and more rows, the combined sizes of all the indicator vectors Ij may be large. However, most elements of a typical indicator vector Ij are zeros, so these vectors can be highly compressed with arithmetic coding. Also, the first few indicator vectors are normally all zeros, so they may all be replaced by one number indicating how many of them there are.  Exercise 7.55: Even though we haven’t yet described the details of the method (specifically, the projection operation), it is possible to logically deduce (or guess) the answer to this exercise. The question is We know that images with little or no correlation will not compress. Yet the size of the last matrix, Mk+1, is small (64 bits or fewer) regardless of the image being compressed. Can quadrisection somehow compress images that other methods cannot? The decoder starts by decoding the prefix, to obtain the sizes of all the compressed strings wj . It then decompresses each wj to obtain an indicator vector Ij . The decoder knows how many indicator vectors there are, since j goes from 1 to k. After decoding all the indicator vectors, the decoder reads the rest of the compressed stream and considers it the bits of Mk+1. The decoder then uses the indicator vectors to reconstruct matrices Mk, Mk-1, and so on all the way down to M0. The projection operation is the heart of quadrisection. It is described here in three steps. Step 1 : An indicator vector I for a matrix M with r rows contains r components, one for each row of M. Each distinct row of M has a zero component in I, and each duplicate row has a positive component. If, for example, row 8 is a duplicate of the 5th distinct row, then the 8th component of I will be 5. Notice that the 5th distinct row is not necessarily row 5. As an example, the indicator vector for matrix M of Equation (7.54) is I = (0, 0, 0, 3, 0, 4, 3, 1): M3 =..... 1001 1101 1011 1011 0000 0000 1011 1001 ..... . (7.54) Step 2 : Given a row vector v of length m, where m is a power of 2, we perform the following: (1) Divide it into .m segments of length .m each. (2) Arrange the segments in a matrix of size .mΧ.m. (3) Divide it into four quadrants, each a matrix of size .m/2Χ.m/2. (4) Label the quadrants 1, 2, 3, and 4 according to 1 2 3 4 . (5) Unfold each into a vector vi of length .m. As an example, consider the vector v = (0, 1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15), of length 16 = 24. It first becomes the 4Χ4 matrix M =... 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 ... . 678 7. Image Compression Partitioning M yields the four 2Χ2 matrices M1 = 0 1 4 5 , M2 = 2 3 6 7 , M3 = 8 9 12 13 , and M4 = 10 11 14 15 . Each is unfolded, to produce the four vectors v1 = (0, 1, 4, 5), v2 = (2, 3, 6, 7), v3 = (8, 9, 12, 13), and v4 = (10, 11, 14, 15). This step illustrates the relation between quadrisection and quadtrees. Step 3 : This finally describes the projection operation. Given an nΧm matrix M where m (the number of columns) is a power of 2, we first construct an indicator vector I for it. Vector I will have n components, for the n rows of M. If M has r distinct rows, vector I will have r zeros corresponding to these rows. We construct the projected matrix M with 4r rows and m/4 columns in the following steps: (3.1) Ignore all duplicate rows of M (i.e., all rows corresponding to nonzero elements of I). (3.2) For each of the remaining r distinct rows of M construct four vectors vi as shown in Step 2 above, and make them into the next four rows of M. We use the notation M I› M, to indicate this projection. Example: Given the 16Χ16 matrix M0 of Figure 7.173 we concatenate its rows to construct matrix M1 with one row and 256 columns. This is always our starting point, regardless of whether the rows of M0 are distinct or not. Since M1 has just one row, its indicator vector is I1 = (0). 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 1 0 0 0 0 0 0 0 0 0 0 0 0 0 1 1 1 1 0 0 0 0 0 0 0 0 0 0 0 0 1 1 0 1 1 1 0 0 0 0 0 0 0 0 0 0 1 1 1 1 1 0 0 0 0 0 0 0 0 0 1 1 1 1 1 1 0 0 0 0 0 0 0 0 1 1 1 1 1 1 1 1 0 0 0 0 0 0 0 1 1 1 1 1 1 1 1 1 0 0 0 0 0 0 0 1 1 1 1 1 1 1 1 0 0 0 0 0 0 0 1 1 1 1 1 1 1 1 1 0 0 0 0 0 0 0 1 1 1 1 1 1 1 1 0 0 0 0 0 0 0 0 1 1 1 1 1 0 1 1 1 0 0 0 0 0 0 0 1 1 1 1 1 0 0 1 1 0 0 0 0 0 0 0 1 1 1 1 1 1 0 1 1 0 0 0 0 0 0 0 0 1 1 1 1 1 0 1 1 0 0 0 0 0 0 0 0 1 1 1 1 1 0 1 1 1 0 0 1 1 1 1 Figure 7.173: A 16Χ16 Matrix M0. 7.35 Quadrisection 679 Matrix M2 is easily projected from M1. It has four rows and 64 columns M2 = 0000000000000001000000110000001100000011000011110011111101111111 1000000010000000110000000111000011100000110000001100000011000000 0111111111111111111111111111101111111001111111010111110101111101 1000000010000000000000001000000010000000100000001000000011001111. All four rows of M2 are distinct, so its indicator vector is I2 = (0, 0, 0, 0). Notice how each row of M2 is an 8Χ8 submatrix of M0. The top row, for example, is the upper-left 8Χ8 corner of M0. To project M3 from M2 we perform Step 2 above on each row of M2, converting it from a 1Χ64 to a 4Χ16 matrix. Thus, matrix M3 consists of four parts, each 4Χ16, so its size is 16Χ16: M3 = ............... 0000000000000000 0000000100110011 0000000000110111 0011111111111111 1000100011000111 0000000000000000 1110110011001100 0000000000000000 0111111111111111 1111111111111011 1111111101110111 1001110111011101 1000100000001000 0000000000000000 1000100010001100 0000000000001111 ............... . Notice how each row of M3 is a 4Χ4 submatrix of M0. The top row, for example, is the upper-left 4Χ4 corner of M0, and the second row is the upper-second-fromleft 4Χ4 corner. Examining the 16 rows of M3 results in the indicator vector I3 = (0, 0, 0, 0, 0, 1, 0, 1, 0, 0, 0, 0, 0, 1, 0, 0) (i.e., rows 6, 8, and 14 are duplicates of row 1). This is how the next projection creates some compression. Projecting M3 to M4 is done by ignoring rows 6, 8, and 14. M4 [Equation (7.55)] therefore has 4 Χ 13 = 52 rows and four columns (a quarter of the number of M3). It therefore has 52 Χ 4 = 208 elements instead of 256. It is clear that with so many short rows, M4 must have many duplicate rows. An examination indicates only 12 distinct rows, and produces vector I4: I4 = (0, 1, 1, 1, 1, 0, 1, 0, 1, 1, 2, 3, 0, 3, 3, 3|0, 1, 0, 4, 3, 0, 3, 1| 0, 3, 3, 3, 3, 3, 0, 3, 3, 3, 0, 3, 0, 10, 3, 10|5, 1, 0, 1, 5, 1, 11, 1, 1, 1, 4, 4). Projecting M4 to obtain M5 is done as before. Matrix M5 has 4Χ12 = 48 rows and only one column (a quarter of the number of columns of M4). This is therefore the final matrix Mk+1, whose transpose will be the last part of the compressed stream.  Exercise 7.56: Compute matrix M5. The compressed stream consists therefore of a prefix, followed by the value of k (= 4), followed by I1, I2, I3, and I4, arithmetically encoded, followed by the 48 bits of MT 5 . There is no need to encode those last bits, since there may be at most 64 of them (see exercise 7.57). Moreover, since I1 and I2 are zeros, there is no need to write them on the output. All that the encoder has to write instead is the number 3 (encoded) to point out that the first indicator vector included in the output is I3. 680 7. Image Compression M4 = ......................................................... 0000 0000 0000 0000 0000 0001 0000 1111 0000 0000 0001 1111 0011 1111 1111 1111 1010 0000 1101 0011 1111 1000 1111 0000 0111 1111 1111 1111 1111 1111 1110 1111 1111 1111 0101 1111 1011 0101 1111 0101 1010 0000 0010 0000 1010 0000 1011 0000 0000 0000 0011 0011 ......................................................... . (7.55) The prefix consists of 4 (k), 3 (index of first nonzero indicator vector), and the encoded lengths of indicator vectors I3 and I4. After decoding this, the decoder can decode the two indicator vectors, read the remaining compressed stream as MT 5 , then use all this data to reconstruct M4, M3, M2, M1, and M0.  Exercise 7.57: Show why M5 cannot have more than 64 elements. Extensions: Quadrisection is used to compress two-dimensional data, and can therefore be considered the second method in a succession of three compression methods the first and third of which are bisection and octasection. Following is a short description of these two methods. Bisection is an extension (reduction) of quadrisection for the case of one-dimensional data (typical examples are sampled audio, a binary string, or text). We assume that the data consists of a string L0 of 2k data items, where an item can be a single bit, an ASCII code, a audio sample, or anything else, but all items have the same size. 7.35 Quadrisection 681 The encoder iterates k times, varying j from 1 to k. In iteration j, a list Lj is created by bisecting the elements of the preceding list Lj-1. An indicator vector Ij is constructed for Lj . The duplicate elements of Lj are then deleted (this is where we get compression). Each of the two elements of the first constructed list L1 is therefore a block of 2k-1 data items (half the number of the data items in L0). Indicator vector I1 also has two elements. Each element of L2 is a block of 2k-2 data items, but the number of elements of L2 is not necessarily four. It can be smaller depending on how many distinct elements L1 has. In the last step, where j = k, list Lk is created, where each element is a block of size 2k-k = 1. Thus, each element of Lk is one of the original data items of L0. It is not necessary to construct indicator vector Ik (the last indicator vector generated is Ik-1). The compressed stream consists of k, followed by indicator vectors I1, I2, through Ik-1 (compressed), followed by Lk (raw). Since the first few indicator vectors tend to be all zeros, they can be replaced by a single number indicating the index of the first nonzero indicator vector. Example: The 32-element string L0 where each data element is a single bit: L0 = (0, 1, 1, 1, 0, 0, 1, 0, 1, 1, 0, 0, 0, 0, 1, 0, 0, 1, 1, 0, 1, 1, 1, 1, 0, 1, 0, 1, 1, 0, 1, 0). The two halves of L0 are distinct, so L1 consists of two elements L1 = (0111001011000010, 0110111101011010), and the first indicator vector is I1 = (0, 0). The two elements of L1 are distinct, so L2 has four elements, L2 = (01110010, 11000010, 01101111, 01011010), and the second indicator vector is I2 = (0, 0, 0, 0). The four elements of L2 are distinct, so L3 has eight elements, L3 = (0111, 0010, 1100, 0010, 0110, 1111, 0101, 1010), and the third indicator vector is I3 = (0, 0, 0, 2, 0, 0, 0, 0). Seven elements of L3 are distinct, so L4 has 14 elements, L4 = (01, 11, 00, 10, 11, 00, 01, 10, 11, 11, 01, 01, 10, 10), and the fourth indicator vector is I4 = (0, 0, 0, 0, 2, 3, 1, 4, 2, 2, 1, 1, 4, 4). Only four elements of L4 are distinct, so L5 has eight elements, L5 = (0, 1, 1, 1, 0, 0, 1, 0). The output therefore consists of k = 5, indicator vectors I1 = (0, 0), I2 = (0, 0, 0, 0), I3 = (0, 0, 0, 2, 0, 0, 0, 0), and I4 = (0, 0, 0, 0, 2, 3, 1, 4, 2, 2, 1, 1, 4, 4) (encoded), followed by L5 = (0, 1, 1, 1, 0, 0, 1, 0). Since the first nonzero indicator vector is I3, we can omit both I1 and I2 and replace them with the integer 3. 682 7. Image Compression  Exercise 7.58: Describe the operations of the decoder for this example.  Exercise 7.59: Use bisection to encode the 32-bit string L0 = (0101010101010101 1010101010101010). The discussion above shows that if the original data to be compressed is a bit string of length 2k, then Lk-1 is a list of pairs of bits. Lk-1 can therefore have at most four distinct elements, so Lk, whose elements are single bits, can have at most eight elements. This, of course, does not mean that any binary string L0 can be compressed into eight bits. The reader should bear in mind that the compressed stream must also include the nonzero indicator vectors. If the elements of L0 are not bits, then Lk could, in principle, be as long as L0. Example: A source string L0 = (a1, a2, . . . , a32) where the ai’s are distinct data items, such as ASCII characters. L1 consists of two elements, L1 = (a1a2 . . . a16, a17a18 . . . a32), and the first indicator vector is I1 = (0, 0). The two elements of L1 are distinct, so L2 has four elements, L2 = (a1a2 . . . a8, a9a10 . . . a16, a17a18 . . . a24, a25a26 . . . a32), and the second indicator vector is I2 = (0, 0, 0, 0). All four elements of L2 are distinct, so L3 = (a1a2a3a4, a5a6a7a8, a9a10a11a12, . . . , a29a30a31a32). Continuing this way, it is easy to see that all indicator vectors will be zero, and L5 will have the same 32 elements as L0. The result will be no compression at all. If the length L of the original data is not a power of two, we can still use bisection by considering the following: There is some integer k such that 2k-1 < L < 2k. If L is close to 2k, we can add d = 2k - L zeros to the original string L0, compress it by bisection, and write d on the compressed stream, so the decoder can delete the d zeros. If L is close to 2k-1 we divide L0 into a string L10 with the first 2k-1 items, and string L20 with the remaining items. The former string can be compressed with bisection and the latter can be compressed by either appending zeros to it or splitting it recursively. Octasection is an extension of both bisection and quadrisection for three-dimensional data. Examples of such data are a video (a sequence of images), a grayscale image, and a color image. Such an image can be viewed as a three-dimensional matrix where the third dimension is the bits constituting each pixel (the bitplanes). We assume that the data is a rectangular box P of dimensions 2k1Χ2k2Χ2k3 , where each entry is a data item (a single bit or several bits) and all entries have the same size. The encoder performs k iterations, where k = min(k1, k2, k3), varying j from 1 to k. In iteration j, a list Lj is created, by subdividing each element of the preceding list Lj-1 into eight smaller rectangular boxes. An indicator vector Ij is constructed for Lj . The duplicate elements of Lj are then deleted (this is where we get compression). 7.36 Space-Filling Curves 683 The compressed stream consists, as before, of all the (arithmetically encoded) indicator vectors, followed by the last list Lk. 7.36 Space-Filling Curves A space-filling curve [Sagan 94] is a parametric function P(t) that passes through every point in a given two-dimensional area, normally the unit square, when its parameter t varies in the range [0, 1]. For any value t0, the value of P(t0) is a point [x0, y0] in the unit square. Mathematically, such a curve is a mapping from the interval [0, 1] to the two-dimensional interval [0, 1] Χ [0, 1]. To understand how such a curve is constructed it is best to think of it as the limit of an infinite sequence of curves P1(t), P2(t), . . ., which are drawn inside the unit square, where each curve is derived from its predecessor by a process of refinement. The details of the refinement depend on the specific curve. Section 7.37.1 discusses several well-known space-filling curves, among them the Hilbert curve and the Sierpi΄nski curve. Since the sequence of curves is infinite, it is impossible to compute all its components. Fortunately, we are interested in a curve that passes through every pixel in a bitmap, not through every mathematical point in the unit square. Since the number of pixels is finite, it is possible to construct such a curve in practice. To understand why such curves are useful for image compression, the reader should recall the principle that has been mentioned several times in the past, namely, if we select a pixel in an image at random, there is a good chance that its neighbors will have the same (or similar) colors. Both RLE image compression and the quadtree method are based on this principle, but they are not always efficient, as Figure 7.174 shows. This 8Χ8 bitmap has two concentrations of pixels, but neither RLE nor the quadtree method compress it very well, since there are no long runs and since the pixel concentrations happen to cross quadrant boundaries. Figure 7.174: An 8Χ8 Bitmap. Better compression may be produced by a method that scans the bitmap area by area instead of line by line or quadrant by quadrant. This is why space-filling curves provide a new approach to image compression. Such a curve visits every point in a given area, and does that by visiting all the points in a subarea, then moves to the next subarea and traverses it, and so on. We use the Hilbert curve (Section 7.37.1) as an example. Each curve Hi is constructed by making four copies of the previous curve 684 7. Image Compression Hi-1, shrinking them, rotating them, and connecting them. The new curve ends up covering the same area as its predecessor. This is the refinement process for the Hilbert curve. Scanning an 8 Χ 8 bitmap in a Hilbert curve results in the sequence of pixels (0,0), (0,1), (1,1), (1,0), (2,0), (3,0), (3,1), (2,1), (2,2), (3,2), (3,3), (2,3), (1,3), (1,2), (0,2), (0,3), (0,4), (1,4), (1,5), (0,5), (0,6), (0,7), (1,7), (1,6), (2,6), (2,7), (3,7), (3,6), (3,5), (2,5), (2,4), (3,4), (4,4), (5,4), (5,5), (4,5), (4,6), (4,7), (5,7), (5,6), (6,6), (6,7), (7,7), (7,6), (7,5), (6,5), (6,4), (7,4), (7,3), (7,2), (6,2), (6,3), (5,3), (4,3), (4,2), (5,2), (5,1), (4,1), (4,0), (5,0), (6,0), (6,1), (7,1), (7,0). Section 7.37.1 discusses space-filling curves in general and shows methods for fast traversal of some curves. Here we would like to point out that quadtrees (Section 7.34) are a special case of the Hilbert curve, a fact illustrated by six figures in that Section. See also [Prusinkiewicz and Lindenmayer 90] and [Prusinkiewicz et al. 91].  Exercise 7.60: Scan the 8Χ8 bitmap of Figure 7.174 using a Hilbert curve and calculate the runs of identical pixels and compare them to the runs produced by RLE. There are actually two types of the Hilbert space-filling curve. Relations between these types, Lindemayer’s L-systems, and Gray codes can be found in Hilbert Curve by Eric Weisstein, available from MathWorld—A Wolfram Web Resource http://mathworld.wolfram.com/HilbertCurve.html 7.37 Hilbert Scan and VQ The space-filling Hilbert curve (Section 7.37.1) is employed by this method [Sampath and Ansari 93] as a preprocessing step for the lossy compression of images (notice that the authors use the term “Peano scan,” but they actually use the Hilbert curve). The Hilbert curve is used to transform the original image into another, highly correlated, image. The compression itself is done by a vector quantization algorithm that exploits the correlation produced by the preprocessing step. In the preprocessing step, an original image of size MΧN is partitioned into small blocks of mΧm pixels each (with m = 8 typically) that are scanned in Hilbert curve order. The result is a linear (one-dimensional) sequence of blocks, which is then rearranged into a new two-dimensional array of size M m ΧNm. This process results in bringing together (or clustering) blocks that are highly correlated. The “distance” between adjacent blocks is now smaller than the distance between blocks that are raster-scan neighbors in the original image. The distance between two blocks Bij and Cij of pixels is measured by the mean absolute difference, a quantity defined by 1 m2 m i=1 m j=1 |Bij - Cij |. 7.37 Hilbert Scan and VQ 685 It turns out that the mean absolute difference of adjacent blocks in the new, rearranged, image is about half that of adjacent blocks in a raster scan of the original image. The reason for this is that the Hilbert curve is space filling. Points that are nearby on the curve are nearby in the image. Conversely, points that are nearby in the image are normally nearby on the curve. Thus, the Hilbert curve acts as an image transform. This fact is the main innovation of the method described here. As an example of a Hilbert scan, imagine an image of size MΧN = 128Χ128. If we select m = 8, we get 16Χ16 = 256 blocks that are scanned and made into a onedimensional array. After rearranging, we end up with a new image of size 2Χ128 blocks or 128 8 Χ128·8 = 16Χ1024 pixels. Figure 7.175 shows how the blocks are Hilbert scanned to produce the sequence shown in Figure 7.176. This sequence is then rearranged. The top row of the rearranged image contains blocks 1, 17, 18, 2, 3, 4, . . ., and the bottom row contains blocks 138, 154, 153, 169, 185, 186, . . .. The new, rearranged image is now partitioned into new blocks of 4 Χ 4 = 16 pixels each. The 16 pixels of a block constitute a vector. (Notice that the vector has 16 components, each of which can be one or more bits, depending on the size of a pixel.) The LBG algorithm (Section 7.19) is used for the vector quantization of those vectors. This algorithm calls for an initial codebook, so the implementors chose five images, scanned them at a resolution of 256Χ256, and used them as training images, to generate three codebooks, consisting of 128, 512, and 1024 codevectors, respectively. The main feature of the particular vector quantization algorithm used here is that the Hilbert scan results in adjacent blocks that are highly correlated. As a result, the LBG algorithm frequently assigns the same codevector to a run of blocks, and this fact can be used to highly compress the image. Experiments indicate that as many as 2–10 consecutive blocks (for images with details) and 30–70 consecutive blocks (for images with high spatial redundancy) may participate in such a run. Therefore, the method precedes each codevector with a code indicating the length of the run (one block or several). There are two versions of the method, one with a fixed-length code and the other with a variable-length prefix code. When a fixed-length code is used preceding each codevector, the main question is the size of the code. A long code, such as six bits, allows for runs of up to 64 consecutive blocks with the same codevector. On the other hand, if the image has small details, the runs would be shorter and some of the 6 bits would be wasted. A short code, such as two bits, allows for runs of up to four blocks only, but fewer bits are wasted in the case of a highly detailed image. A good solution is to write the size of the code (which is typically 2–6 bits) at the start of the compressed stream, so the decoder knows what it is. The encoder can then be very sophisticated, trying various code sizes before settling on one and using it to compress the image. Using prefix codes can result in slightly better compression, especially if the encoder can perform a two-pass job to determine the frequency of each run before anything is compressed. In such a case, the best Huffman codes can be assigned to the runs, resulting in best compression. Further improvement can be achieved by a variant of the method that uses dynamic codebook partitioning. This is again based on adjacent blocks being very similar. Even if such blocks end up with different codevectors, those codevectors may be very similar. This variant selects the codevectors for the first block in the usual way, using the entire 686 7. Image Compression 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 69 70 71 72 73 74 75 76 77 78 79 80 81 82 83 84 85 86 87 88 89 90 91 92 93 94 95 96 97 98 99 100 101 102 103 104 105 106 107 108 109 110 111 112 113 114 115 116 117 118 119 120 121 122 123 124 125 126 127 128 129 130 131 132 133 134 135 136 137 138 139 140 141 142 143 144 145 146 147 148 149 150 151 152 153 154 155 156 157 158 159 160 161 162 163 164 165 166 167 168 169 170 171 172 173 174 175 176 177 178 179 180 181 182 183 184 185 186 187 188 189 190 191 192 192 194 195 196 197 198 199 200 201 202 203 204 205 206 207 208 209 210 211 212 213 214 215 216 217 218 219 220 221 222 223 224 225 226 227 228 229 230 231 232 233 234 235 236 237 238 239 240 241 242 243 244 245 246 247 248 249 250 251 252 253 254 255 256 1 Figure 7.175: Hilbert Scan of 16Χ16 Blocks. 1, 17, 18, 2, 3, 4, 20, 19, 35, 36, 52, 51, 50, 34, 33, 49, 65, 66, 82, 81, 97, 113, 114, 98, 99, 115, 116, 100, 84, 83, 67, 68, 69, 70, 86, 85, 101, 117, 118, 102, 103, 119, 120, 104, 88, 87, 71, 72, 56, 40, 39, 55, 54, 53, 37, 38, 22, 21, 5, 6, 7, 23, 24, 8, 9, 10, 26, 25, 41, 57, 58, 42, 43, 59, 60, 44, 28, 27, 11, 12, 13, 29, 30, 14, 15, 16, 32, 31, 47, 48, 64, 63, 62, 46, 45, 61, 77, 93, 94, 78, 79, 80, 96, 95, 111, 112, 128, 127, 126, 110, 109, 125, 124, 123, 107, 108, 92, 76, 75, 91, 90, 74, 73, 89, 105, 106, 122, 121, 137, 138, 154, 153, 169, 185, 186, 170, 171, 187, 188, 172, 156, 155, 139, 140, 141, 157, 158, 142, 143, 144, 160, 159, 175, 176, 192, 191, 190, 174, 173, 189, 205, 221, 222, 206, 207, 208, 224, 223, 239, 240, 256, 255, 254, 238, 237, 253, 252, 251, 235, 236, 220, 204, 203, 219, 218, 202, 201, 217, 233, 234, 250, 249, 248, 232, 231, 247, 246, 245, 229, 230, 214, 213, 197, 198, 199, 215, 216, 200, 184, 183, 167, 168, 152, 136, 135, 151, 150, 134, 133, 149, 165, 166, 182, 181, 180, 179, 163, 164, 148, 132, 131, 147, 146, 130, 129, 145, 161, 162, 178, 177, 192, 209, 210, 194, 195, 196, 212, 211, 227, 228, 244, 243, 242, 226, 225, 241 Figure 7.176: The Resulting 256 Blocks. 7.37 Hilbert Scan and VQ 687 codebook. It then selects a set of the next best codevectors that could have been used to code this block. This set becomes the active part of the codebook. (Notice that the decoder can mimic this selection.) The second block is then compared to the first block and the distance between them measured. If this distance is less than a given threshold, the codevector for the second block is selected from the active set. Since the active set is much smaller than the entire codebook, this leads to much better compression. However, each codevector must be preceded by a bit telling the decoder whether the codevector was selected from the entire codebook or just from the active set. If the distance is greater than the threshold, a codevector for the second block is selected from the entire codebook, and a new active set is chosen, to be used (hopefully) for the third block. If the Hilbert scan really ends up with adjacent blocks that are highly correlated, a large fraction of the blocks are coded from the active sets, thereby considerably improving the compression. A typical codebook size is 128–1024 entries, whereas the size of an active set may be just four codevectors. This reduces the size of a codebook pointer from 7–10 bits to 2 bits. The choice of the threshold for this variant is important. It seems that an adaptive threshold adjustment may work best, but the developers don’t indicate how this may be implemented. 7.37.1 Examples A space-filling curve completely fills up part of space by passing through every point in that part. It does that by changing direction repeatedly. We will only discuss curves that fill up part of the two-dimensional plane, but the concept of a space-filling curve exists for any number of dimensions.  Exercise 7.61: Show an example of a space-filling curve in one dimension. Several such curves are known and all are defined recursively. A typical definition starts with a simple curve C0, shows how to use it to construct another, more complex curve C1, and defines the final, space-filling curve as the limit of the sequence of curves C0,C1, . . . . 7.37.2 The Hilbert Curve (This discussion is based on the approach of [Wirth 76].) Perhaps the most familiar of these curves is the Hilbert curve, discovered by the great mathematician David Hilbert in 1891. The Hilbert curve [Hilbert 91] is the limit of a sequence H0,H1,H2 . . . of curves, some of which are shown in Figure 7.177. They are defined by the following: 0. H0 is a single point. 1. H1 consists of four copies of (the point) H0, connected with three straight segments of length h at right angles to each other. Four orientations of this curve, labeled 1, 2, 3, and 4, are shown in Figure 7.177a. 2. The next curve, H2, in the sequence is constructed by connecting four copies of different orientations of H1 with three straight segments of length h/2 (shown in bold in Figure 7.177b). Again there are four possible orientations of H2, and the one shown is #2. It is constructed of orientations 1223 of H1, connected by segments that go to the 688 7. Image Compression 2 1 3 4 1 2 3 2 (a) (b) (c) Figure 7.177: Hilbert Curves of Orders 1, 2, and 3. right, up, and to the left. The construction of the four orientations of H2 is summarized in Table Ans.41. Curve H3 is shown in Figure 7.177c. The particular curve shown is orientation 1223 of H2. Figures 7.178, 7.179 and 7.180 show the Hilbert curves of orders 4, 5 and 6. It is easy to see how fast these curves become extremely complex. 7.37.3 The Sierpi΄nski Curve Another well-known space-filling curve is the Sierpi΄nski curve. Figure 7.181 shows curves S1 and S2, and Sierpi΄nski has proved [Sierpi΄nski 12] that the limit of the sequence S1, S2, . . . is a curve that passes through every point of the unit square [0, 1] Χ [0, 1]. To construct this curve, we need to figure out how S2 is constructed out of four copies of S1. The first thing that comes to mind is to follow the construction method used for the Hilbert curve, i.e., to take four copies of S1, eliminate one edge in each, and connect them. This, unfortunately, does not work, because the Sierpi΄nski curve is very different from the Hilbert curve. It is closed, and it has one orientation only. A better approach is to start with four parts that constitute four orientations of one open curve, and connect them with straight segments. The segments are shown dashed in Figure 7.181. Notice how Figure 7.181a is constructed of four orientations of a basic, three-part curve connected by four short, dashed segments. Figure 7.181b is similarly constructed of four orientations of a complex, 15-part curve, connected by the same short, dashed segments. If we denote the four basic curves A, B, C, and D, then the basic construction rule of the Sierpi΄nski curve is S: ABC D , and the recursion rules are: A: AB››D A B: BC v v A B C: C D‹‹BC D: D A ^ ^ C D (7.56) Figure 7.182 shows the five Sierpi΄nski curves of orders 1 through 5 superimposed on each other.  Exercise 7.62: Figure 7.183 shows three iterations of the Peano space-filling curve, developed in 1890. Use the techniques developed earlier for the Hilbert and Sierpi΄nski 7.37 Hilbert Scan and VQ 689 1 4 2 3 2 4 3 2 4 3 2 1 2 4 1 3 3 1 2 2 4 2 4 1 1 3 2 3 3 3 3 2 4 1 2 1 1 1 1 2 4 3 2 2 4 2 4 3 3 1 2 3 2 4 3 1 1 3 2 2 1 4 1 2 Figure 7.178: Hilbert Curve of Order 4. Figure 7.179: Hilbert Curve of Order 5. 690 7. Image Compression Figure 7.180: Hilbert Curve of Order 6. (a) (b) Figure 7.181: Sierpi΄nski Curves of Orders 1 and 2. 7.37 Hilbert Scan and VQ 691 Figure 7.182: Sierpi΄nski Curves of Orders 1–5. curves, to describe how the Peano curve is constructed. (Hint: The curves shown are P1, P2, and P3. The first curve, P0, in this sequence is not shown.) 7.37.4 Traversing the Hilbert Curve Space-filling curves are used in image compression (Section 7.36), which is why it is important to develop methods for a fast traversal of such a curve. Two approaches, both table-driven, are illustrated here for traversing the Hilbert curve. The first approach [Cole 86] is based on the observation that the Hilbert curve Hi is constructed of four copies of its predecessor Hi-1 placed at different orientations. A look 692 7. Image Compression (a) (b) (c) Figure 7.183: Three Iterations of the Peano Curve. at Figures 7.177, 7.178, 7.179, and 7.180 should convince the reader that Hi consists of 22i nodes connected with straight segments. The node numbers therefore vary from 0 to 22i - 1 and require 2i bits each. In order to traverse the curve we need a function that computes the coordinates (x, y) of a node i from the node number i. The (x, y) coordinates of a node in Hi are i-bit numbers. A look at Figure 7.178 shows how successive nodes are initially located at the bottom-left quadrant, and then move to the bottom-right quadrant, the top-right quadrant, and finally the top-left one. This figure shows orientation #2 of the curve, so we can say that this orientation of Hi traverses quadrants 0, 1, 2, and 3, where quadrants are numbered 3 2 0 1. It is now clear that the two leftmost bits of a node number determine its quadrant. Similarly, the next pair of bits in the node number determine its subquadrant within the quadrant, but here we run into the added complication that each subquadrant is placed at a different orientation in its quadrant. This approach therefore uses Table 7.184 to determine the coordinates of a node from its number. Bit Next Bit Next Bit Next Bit Next pair x y table pair x y table pair x y table pair x y table 00 0 0 2 00 0 0 1 00 1 1 4 00 1 1 3 01 1 0 1 01 0 1 2 01 0 1 3 01 1 0 4 10 1 1 1 10 1 1 2 10 0 0 3 10 0 0 4 11 0 1 4 11 1 0 3 11 1 0 2 11 0 1 1 (1) (2) (3) (4) Table 7.184: Coordinates of Nodes in Hi. As an example, we compute the xy coordinates of node 109 (the 110th node) of orientation #2 of H4. The H4 curve has 22·4 = 256 nodes, so node numbers are eight bits each, and 109 = 011011012. We start with Table 7.184(1). The two leftmost bits of the node number are 01, and table (1) tells us that the x coordinate start with 1, the y coordinate, with 0, and we should continue with table (1). The next pair of bits is 10, and table (1) tells us that the next bit of x is 1, the next bit of y is 1, and we should 7.37 Hilbert Scan and VQ 693 stay with table (1). The third pair of bits is 11, so table (1) tells us that the next bit of x is 0, the next bit of y is 1, and we should switch to table (4). The last pair of bits is 01, and table (4) tells us to append 1 and 0 to the coordinates of x and y, respectively. Thus, the coordinates are x = 1101 = 13, y = 0110 = 6, as can be verified directly from Figure 7.178 (the small circle). It is also possible to transform a pair of coordinates (x, y), each in the range [0, 2i-1], to a node number in Hi by means of Table 7.185. xy Int. Next xy Int. Next xy Int. Next xy Int. Next pair pair table pair pair table pair pair table pair pair table 00 00 2 00 00 1 00 10 3 00 10 4 01 11 4 01 01 2 01 01 3 01 11 1 10 01 1 10 11 3 10 11 2 10 01 4 11 10 1 11 10 2 11 00 4 11 00 3 (1) (2) (3) (4) Table 7.185: Node Numbers in Hi.  Exercise 7.63: Use Table 7.185 to compute the node number of the H4 node whose coordinates are (13, 6). The second approach to Hilbert curve traversal uses Table Ans.41. Orientation #2 of the H2 curve shown in Figure 7.177(b) is traversed in order 1223. The same orientation of the H3 curve of Figure 7.177(c) is traversed in 2114 1223 1223 4332, but Table Ans.41 tells us that 2114 is the traversal order for orientation #1 of H2, 1223 is the traversal for orientation #2 of H2, and 4332 is for orientation #3. The traversal of orientation #2 of H3 is therefore also based on the sequence 1223. Similarly, orientation #2 of H4 is traversed (Figure 7.178) in the order 1223 2114 2114 3441 2114 1223 1223 4332 2114 1223 1223 4332 3441 4332 4332 1223, which is reduced to 2114 1223 1223 4332, which in turn is reduced to the same sequence 1223. The idea is therefore to create the traversal order for orientation #2 of Hi by starting with the sequence 1223 and recursively expanding it i - 1 times, using Table Ans.41.  Exercise 7.64: (Easy.) Show how to apply this method to traversing orientation #1 of Hi. A MATLAB function hilbert.m to compute the traversal of the curve is available at [Matlab 99]. It was written by Daniel Leo Lau (dllau@engr.uky.edu). The call hilbert(4) produces the 4 Χ 4 matrix 5 6 9 10 4 7 8 11 3 2 13 12 0 1 14 15 . 694 7. Image Compression 7.37.5 Traversing the Peano Curve The Peano curves P0, P1, and P2 of Figure Ans.40 have 1, 32, and 34 nodes, respectively. In general, Pn has 32n nodes, numbered 0, 1, 2, . . . , 32n-1. This suggests that the Peano curve [Peano 90] is somehow based on the number 3, in contrast with the Hilbert curve, which is based on 2. The coordinates of the nodes vary from (0, 0) to (n - 1, n - 1). It turns out that there is a correspondence between the node numbers and their coordinates [Cole 85], which uses base-3 reflected Gray codes (Section 7.4.1). A reflected Gray code [Gray 53] is a permutation of the i-digit integers such that consecutive integers differ by one digit only. Here is one way to derive these codes for binary numbers. Start with i = 1. There are only two 1-bit digits, namely, 0 and 1, and they differ by 1 bit only. To get the RGC for i = 2 proceed as follows: 1. Copy the sequence (0, 1). 2. Append (on the left or on the right) a 0 bit to the original sequence and a bit of 1 to the copy. The result is (00, 01), (10, 11). 3. Reflect (reverse) the second sequence. The result is (11, 10). 4. Concatenate the two sequences to get (00, 01, 11, 10). It is easy to see that consecutive numbers differ by one bit only.  Exercise 7.65: Follow the rules above to get the binary RGC for i = 3. Notice that the first and last numbers in an RGC also differ by one bit. RGCs can be created for any number system using the following notation and rules: Let a = a1a2 · · · am be a non-negative, base-n integer (i.e., 0 . ai < n). Define the quantity pj = (j i=1 ai) mod 2, and denote the base-n RGC of a by a = b1b2 · · · bm. The digits bi of a can be computed by b1 = a1; bi =.. .... ... ai if pi-1 = 0; Odd n n - 1 - ai if pi-1 = 1. ai if ai-1 is even; Even n n - 1 - ai if ai-1 is odd. i = 2, 3, . . . , m. [Note that (a) = a for both even and odd n.] For example, the RGC of the sequence of base-3 numbers (trits) 000, 001, 002, 010, 011, 012, 020, 021, 022, 100, 101. . . is 000, 001, 002, 012, 011, 010, 020, 021, 022, 122, 121. . . . The connection between Peano curves and RGCs is as follows: Let a be a node in the Peano curve Pm. Write a as a ternary (base-3) number with 2m trits a = a1a2 · · · a2m. Let a = b1b2 · · · b2m be the RGC equivalent of a. Compute the two numbers x = b2b4b6 · · · b2m and y = b1b3b5 · · · b2m-1. The number x is the RGC of a number x, and similarly for y. The two numbers (x, y) are the coordinates of node a in Pm. 7.38 Finite Automata Methods 695 7.38 Finite Automata Methods Finite automata methods are somewhat similar to the IFS method of Section 7.39. They are based on the fact that images used in practice have a certain amount of self-similarity, i.e., it is often possible to find a part of the image that looks the same (or almost the same) as another part, except perhaps for size, brightness, or contrast. It is shown in Section 7.39 that IFS uses affine transformations to match image parts. The methods described here, on the other hand, try to describe a subimage as a weighted sum of other subimages. Two methods are described in this section, weighted finite automata (or WFA) and generalized finite automata (or GFA). Both are due to Karel Culik, the former in collaboration with Jarkko Kari, and the latter with Vladimir Valenta. The term “automata” is used because these methods represent the image to be compressed in terms of a graph that is very similar to graphs used to represent finite-state automata (also called finite-state machines, see Section 11.8). The main references are [Culik and Kari 93], [Culik and Kari 94a,b], and [Culik and Kari 95]. 7.38.1 Weighted Finite Automata WFA starts with an image to be compressed. It locates image parts that are identical or very similar to the entire image or to other parts, and constructs a graph that reflects the relationships between these parts and the entire image. The various components of the graph are then compressed and become the compressed image. The image parts used by WFA are obtained by a quadtree partitioning of the image. Any quadrant, subquadrant, or pixel in the image can be represented by a string of the digits 0, 1, 2, and 3. The longer the string, the smaller the image area (subsquare) it represents. We denote such a string by a1a2 . . . ak. Quadtrees were introduced in Section 7.34. We assume that the quadrant numbering of Figure 7.186a is extended recursively to subquadrants. Figure 7.186b shows how each of the 16 subquadrants produced from the 4 original ones are identified by a 2-digit string of the digits 0, 1, 2, and 3. After another subdivision, each of the resulting subsubquadrants is identified by a 3-digit string, and so on. The black area in Figure 7.186c, for example, is identified by the string 1032. 1 3 0 2 11 13 31 33 10 12 30 32 01 03 21 23 00 02 20 22 1032 (a) (b) (c) Figure 7.186: Quadrant Numbering. 696 7. Image Compression  Exercise 7.66: What string identifies the gray area of Figure 7.186c. Instead of constantly saying “quadrant, subquadrant, or subsubquadrant,” we use the term “subsquare” throughout this section.  Exercise 7.67: (Proposed by Karel Culik.) What is special about the particular quadrant numbering of Figure 7.186a? If the image size is 2n Χ2n, then a single pixel is represented by a string of n digits, and a string a1a2 . . . ak of k digits represents a subsquare of size 2n-k Χ 2n-k pixels. Once this is grasped, it is not hard to see how any type of image, bi-level, grayscale, or color, can be represented by a graph. Mathematically, a graph is a set of nodes (or vertices) connected by edges. A node may contain data, and an edge may have a label. The graphs used in WFA represent finite automata, so the term state is used instead of node or vertex. The WFA encoder starts by generating a graph that represents the image to be compressed. One state in this graph represents the entire image, and each of the other states represents a subimage. The edges show how certain subimages can be expressed as linear combinations of the entire (scaled) image or of other subimages. Since the subimages are generated by a quadtree, they do not overlap. The basic rule for connecting states with edges is; If quadrant a (where a can be 0, 1, 2, or 3) of subimage i is identical to subimage j, construct an edge from state i to state j, and label it a. In an arbitrary image it is not very common to find two identical subimages, so the rule above is extended to; If subsquare a of subimage i can be obtained by multiplying all the pixels of subimage j by a constant w, construct an edge from state i to state j, label it a, and assign it a weight w. We use the notation a(w) or, if w = 1, just a. A sophisticated encoding algorithm may discover that subsquare a of subimage i can be expressed as the weighted sum of subimages j and k with weights u and w, respectively. In such a case, two edges should be constructed. One from i to j, labeled a(u), and another, from i to k, labeled a(w). In general, such an algorithm may discover that a subsquare of subimage i can be expressed as a weighted sum (or a linear combination) of several subimages, with different weights. In such a case, an edge should be constructed for each term in the sum. We denote the set of states of the graph by Q and the number of states by m. Notice that the weights do not have to add up to 1. They can be any real numbers, and can be greater than 1 or negative. Figure 7.187b is the graph of a simple 2 Χ 2 chessboard. It is clear that we can ignore subsquares 1 and 2, since they are all white. The final graph includes states for subsquares that are not all white, and it has two states. State 1 is the entire image, and state 2 is an all black subimage. Since subsquares 0 and 3 are identical to state 2, there are two edges (shown as one edge) from state 1 to state 2, where the notation 0, 3(1) stands for 0(1) and 3(1). Since state 2 represents subsquares 0 and 3, it can be named q0 or q3. Figure 7.187c shows the graph of a 2Χ2 chessboard where quadrant 3 is 50% gray. This is also a two-state graph. State 2 of this graph is an all-black image and is identical to quadrant 0 of the image (we can also name it q0). This is why there is an edge labeled 0(1). State 2 can also generate quadrant 3, if multiplied by 0.5, and this is expressed by the edge 3(0.5). Figure 7.187d shows another two-state graph representing the same 7.38 Finite Automata Methods 697 (a) 3(0.5) 3(1.5) 0(1) 0(1) 1 0,3(1) 0,1,2,3(1) (b) (c) (d) 3(1) 0(0.5) 1,2(0.5) 1,2(0.25) 0(-0.25) 1,2(-0.375) 3(-0.5) 1,2(1.25) 0,1,2,3(1) 3(1) 0(2) 0,1,2,3(1) 3 0 2 50 100 50 75 100 0 50 75 1 3 0 2 1 3 0 2 2 2 2 2 1 1 1 1 Figure 7.187: Graphs for Simple Images. image. This time, state 2 (which may be named q3) is a 50%-gray image, and the weights are different. Figure 7.187a shows an example of a weighted sum. The image (state 1) varies smoothly from black in the top-right corner to white in the bottom-left corner. The graph contains one more state, state 2, which is identical to quadrant 3 (we can also name it q3). It varies from black in the top-right corner to 50% gray in the bottom-left corner, with 75% gray in the other two corners. Quadrant 0 of the image is obtained when the entire image is multiplied by 0.5, so there is an edge from state 1 to itself labeled 0(0.5). Quadrants 1 and 2 are identical and are obtained as a weighted sum of the entire image (state 1) multiplied by 0.25 and of state 2 (quadrant 3) multiplied by 0.5. The four quadrants of state 2 are similarly obtained as weighted sums of states 1 and 2. The average grayness of each state is easy to figure out in this simple case. For example, the average grayness of state 1 of Figure 7.187a is 0.5 and that of quadrant 3 of state 2 is (1 + 0.75)/2 = 0.875. The average grayness is used later and is called the final distribution. Notice that the two states of Figure 7.187a have the same size. In fact, nothing has been said so far about the sizes and resolutions of the image and the various subimages. The process of constructing the graph does not depend on the resolution, which is why WFA can be applied to multiresolution images, i.e., images that may exist in several different resolutions. Figure 7.188 is another example. The original image is state 1 of the graph, and the graph has four states. State 2 is quadrant 0 of the image (so it may be called q0). State 3 represents (the identical) quadrants 1 and 2 of the image (it may be called q1 or q2). It should be noted that, in principle, a state of a WFA graph does not have to correspond to any subsquare of the image. However, the recursive inference algorithm used by WFA generates a graph where each state corresponds to a subsquare of the image.  Exercise 7.68: In Figure 7.188, state 2 of the graph is represented in terms of itself and of state 4. Show how to represent it in terms of itself and an all-black state. 698 7. Image Compression 100 100 100 100 100 100 75 82.5 75 75 50 50 50 0 0 50 50 4 3 4 4 1 1 2 2 2 2 3 3 4 0(1) 1,2(1) 0(0.5) 1,2(0.5) 3(1) 1,2(0.25) 1,2(-0.5) 0(1) 0(-0.25) 1,2(1.25) 3(1.5) 1,2(-0.375)3(-1) 0(1) 1,2(1.5) Figure 7.188: A Four-State Graph. A multiresolution imageM may, in principle, be represented by an intensity function fM(x, y) that gives the intensity of the image at point (x, y). For simplicity, we assume that x and y vary independently in the range [0, 1], that the intensity varies between 0 (white) and 1 (black), and that the bottom-left corner of the image is located at the origin [i.e., it is point (0, 0)]. Thus, the subimage of state 2 of Figure 7.188 is defined by the function f(x, y) = x + y 2 .  Exercise 7.69: What function describes the image of state 1 of Figure 7.188? Since we are dealing with multiresolution images, the intensity function f must be able to generate the image at any resolution. If we want, for example, to lower the resolution to one-fourth the original one, each new pixel should be computed from a set of four original ones. The obvious value for such a “fat” pixel is the average of the four original pixels. This implies that at low resolutions, the intensity function f should compute the average intensity of a region in the original image. In general, the value of f for a subsquare w should satisfy f(w) = 1 4 [f(w0) + f(w1) + f(w2) + f(w3)] . Such a function is called average preserving (ap). It is also possible to increase the resolution, but the details created in such a case would be artificial. Given an arbitrary image, however, we generally do not know its intensity function. Also, any given image has limited, finite resolution. Thus, WFA proceeds as follows: It starts with a given image with resolution 2nΧ2n. It uses an inference algorithm to construct its graph, then extracts enough information from the graph to be able to 7.38 Finite Automata Methods 699 reconstruct the image at its original or any lower resolution. The information obtained from the graph is compressed and becomes the compressed image. The first step in extracting information from the graph is to construct four transition matrices W0, W1, W2, and W3 according to the following rule: If there is an edge labeled q from state i to state j, then element (i, j) of transition matrix Wq is set to the weight w of the edge. Otherwise, (Wq)i,j is set to zero. An example is the four transition matrices resulting from the graph of Figure 7.189. They are W0 = 0.5 0 0 1 , W1 = 0.5 0.25 0 1 , W2 = 0.5 0.25 0 1 , W3 = 0.5 0.5 0 1 . 50 100 0 50 1 1,2(0.25) 0,1,2,3(0.5) 0,1,2,3(1) 3(0.5) 2 Figure 7.189: A Two-State Graph. The second step is to construct a column vector F of size m called the final distribution (this is not the same as the intensity function f). Each component of F is the average intensity of the subimage associated with one state of the graph. Thus, for the two-state graph of Figure 7.189 we get F = (0.5, 1)T .  Exercise 7.70: Write the four transition matrices and the final distribution for the graph of Figure 7.188. In the third step we define quantities .i(a1a2 . . . ak). As a reminder, if the size of the original image is 2nΧ2n, then the string a1a2 . . . ak (where ai = 0, 1, 2, or 3) defines a subsquare of size 2n-kΧ2n-k pixels. The definition is .i(a1a2 . . . ak) = (Wa1 ·Wa2 · · ·Wak ·F)i . (7.57) Thus, .i(a1a2 . . . ak) is the ith element of the column (Wa1 ·Wa2 · · ·Wak ·F)T; it is a number. Each transition matrix Wa has dimensions m Χ m (where m the number of states of the graph) and F is a column of size m. For any given string a1a2 . . . ak there are therefore m numbers .i(a1a2 . . . ak), for i = 1, 2, . . . , m. We use the two-state graph of Figure 7.189 to calculate some .i’s: .i(0) = (W0 ·F)i = 0.5 0 0 1 0.5 1 i = 0.25 1 i , 700 7. Image Compression .i(01) = (W0 ·W1 ·F)i = 0.5 0 0 1 0.5 0.25 0 1 0.5 1 i = 0.25 1 i , .i(1) = (W1 ·F)i = 0.5 0.25 0 1 0.5 1 i = 0.5 1 i , (7.58) .i(00) = (W0 ·W0 ·F)i = 0.5 0 0 1 0.5 0 0 1 0.5 1 i = 0.125 1 i , .i(03) = (W0 ·W3 ·F)i = 0.5 0 0 1 0.5 0.5 0 1 0.5 1 i = 0.375 1 i , .i(33 . . . 3) = (W3 ·W3 · · ·W3 ·F)i = 0.5 0.5 0 1 0.5 0.5 0 1 · · · 0.5 0.5 0 1 0.5 1 i = 0 1 0 1 0.5 1 i = 11 i .  Exercise 7.71: Compute .i at the center of the image. Notice that each .i(w) is identified by two quantities, its subscript i (which corresponds to a state of the graph) and a string w = a1a2 . . . ak that specifies a subsquare of the image. The subsquare can be as big as the entire image or as small as a single pixel. The quantity .i (where no subsquare is specified) is called the image of state i of the graph. The name image makes sense, since for each subsquare w, .i(w) is a number, so .i consists of several numbers from which the pixels of the image of state i can be computed and displayed. The WFA encoding algorithms described in this section generate a graph where each state is a subsquare of the image. In principle, however, some states of the graph may not correspond to any subsquare of the image. An example is state 2 of the graph of Figure 7.189. It is all black, even though the image does not have an all-black subsquare. Figure 7.190 shows the image of Figure 7.189 at resolutions 2Χ2, 4Χ4, and 256Χ256. 25 50 75 13 25 37 50 63 75 88 Figure 7.190: Image f = (i + j)/2 at Three Resolutions. 7.38 Finite Automata Methods 701  Exercise 7.72: Use mathematical software to compute a matrix such as those of Figure 7.190. We now introduce the initial distribution I = (I1, I2, . . . , Im), a row vector with m elements. If we set it to I = (1, 0, 0, . . . , 0), then the dot product I·.(a1a2 . . . ak) [where the notation .(q) stands for a vector with the m values .1(q) through .m(q)] gives the average intensity f(a1a2 . . . ak) of the subsquare specified by a1a2 . . . ak. This intensity function can also be expressed as the matrix product f(a1a2 . . . ak) = I ·Wa1 ·Wa2 · · ·Wak ·F. (7.59) In general, the initial distribution specifies the image defined by a WFA as a linear combination I1.1 + · · · + In.n of “state images.” If I = (1, 0, 0, . . . , 0), then the image defined is .1, the image corresponding to the first state. This is always the case for the image resulting from the inference algorithm described later in this section. Given a WFA, we can easily find another WFA for the image obtained by zooming to the subsquare with address a1a2 . . . ak. We just replace the initial distribution I by I ·Wa1 ·Wa2 · · ·Wak. To prove this, consider the subsquare with address b1b2 . . . bm in the zoomed square. It corresponds to the subsquare in the entire image with address a1a2 . . . akb1b2 . . . bm. Hence, the grayness value computed for it by the original WFA is I Wa1Wa2 . . . WakWb1Wb2 . . . Wbm F. Using the new WFA for the corresponding subsquare, we get the same value, namely IWb1Wb2 . . . Wbm F, where I = I Wa1Wa2 . . . Wak (proof provided by Karel Culik). For the .i’s computed in Equation (7.58) the dot products of the form I·.(a1a2 . . . ak) yield f(0) = I ·.(0) = (1, 0) 0.25 1 = I1.1(0) + I2.2(0) = 0.25, f(01) = I ·.(01) = 1Χ0.25 + 0Χ1 = 0.25, f(1) = I ·.(1) = 0.5, f(00) = I ·.(00) = 0.125, f(03) = I ·.(03) = 0.375, f(33 . . . 3) = I ·.(33 . . . 3) = 1.  Exercise 7.73: Compute the .i’s and the corresponding f values for subsquares 0, 01, 1, 00, 03, and 3 of the five-state graph of Figure 7.188. Equation (7.57) is the definition of .i(a1a2 . . . ak). It shows that this quantity is the ith element of the column vector (Wa1 ·Wa2 · · ·Wak ·F)T . We now examine the reduced column vector (Wa2 · · ·Wak ·F)T . Its ith element is, according to the definition of .i, the quantity .i(a2 . . . ak). Thus, we conclude that .i(a1a2 . . . ak) = (Wa1)i,1.1(a2 . . . ak) + (Wa1)i,2.2(a2 . . . ak) + · · · + (Wa1)i,m.m(a2 . . . ak), 702 7. Image Compression or, if we denote the string a2 . . . ak by w, .i(a1w) = (Wa1)i,1.1(w) + (Wa1)i,2.2(w) + · · · + (Wa1)i,m.m(w) = m j=1 (Wa1)i,j.j(w). (7.60) The quantity .i(a1w) (which corresponds to quadrant a1 of subsquare w) can be expressed as a linear combination of the quantities .j(w), where j = 1, 2, . . . , m. This justifies calling .i the image of state i of the graph. We say that subsquare a1 of image .i(w) can be expressed as a linear combination of images .j(w), for j = 1, 2, . . . , m. This is how the self-similarity of the original image enters the picture. Equation (7.60) is recursive, since it defines a smaller image in terms of larger ones. The largest image is, of course, the original image, for which w is null (we denote the null string by ). This is where the recursion starts .i(a) = (Wa)i,1.1() + (Wa)i,2.2() + · · · + (Wa)i,m.m() = m j=1 (Wa)i,j.j(), for a = 0, 1, 2, 3. (7.61) On the other hand, Equation (7.57) shows that .i() = Fi, so Equation (7.61) becomes .i(a) = m j=1 (Wa)i,jFj , for a = 0, 1, 2, 3. (7.62) It should now be clear that if we know the four transition matrices and the final distribution, we can compute the images .i(w) for strings w of any length (i.e., for subsquares of any size). Once an image .i(w) is known, the average intensity f(w) of subsquare w can be computed by Equation (7.59) (which requires knowledge of the initial distribution). The problem of representing an image f in WFA is thus reduced to finding vectors I and F and matrices W0, W1, W2, and W3 that will produce f (or an image close to it). Alternatively, we can find I and images .i() for i = 1, 2, . . . , m. Here is an informal approach to this problem. We can start with a graph consisting of one state and select this state as the entire image, .1(). We now concentrate on quadrant 0 of the image and try to determine .1(0). If quadrant 0 is identical, or similar enough, to (a scaled version of) the entire image, we can write .1(0) = .1() = m j=1 (W0)1,j.j(), which is true if the first row of W0 is (1, 0, 0, . . . , 0). We have determined the first row of W0, even though we don’t yet know its size (it is m but m hasn’t been determined 7.38 Finite Automata Methods 703 yet). If quadrant 0 is substantially different from the entire image, we add .1(0) to the graph as a new state, state 2 (i.e., we increment m by 1) and call it .2(0). The result is .1(0) = .2(0) = m j=1 (W0)1,j.j(0), which is true if the first row of W0 is (0, 1, 0, . . . , 0). We have again determined the first row of W0, even though we still don’t know its size. Next, we process the remaining three quadrants of .1() and the four quadrants of .2(0). Let’s examine, for example, the processing of quadrant three [.2(03)] of .2(0). If we can express it (precisely or close enough) as the linear combination .2(03) = ..1(3) + ί.2(3) = j (W0)2,j.j(3), then we know that the second row of W0 must be (., ί, 0, . . . , 0). If we cannot express .2(03) as a linear combination of already known .’s, then we declare it a new state .3(03), and this implies (W3)2,3 = 1, and also determines the second row of W0 to be (0, 0, 1, 0, . . . , 0). This process continues until all quadrants of all the .j ’s have been processed. This normally generates many more .i’s, all of which are subsquares of the image. This intuitive algorithm is now described more precisely, using pseudocode. It constructs a graph from a given multiresolution image f one state at a time. The graph has the minimal number of states (to be denoted by m) but a relatively large number of edges. Since m is minimal, the four transition matrices (whose dimensions are mΧm) are small. However, since the number of edges is large, most of the elements of these matrices are nonzero. The image is compressed by writing the transition matrices (in compressed format) on the compressed stream, so the sparser the matrices, the better the compression. Thus, this algorithm does not produce good compression and is described here because of its simplicity. We denote by i the index of the first unprocessed state, and by ., a mapping from states to subsquares. The steps of this algorithm are as follows: Step 1: Set m = 1, i = 1, F(q1) = f(), .(q1) = . Step 2: Process qi, i.e., for w = .(qi) and a = 0, 1, 2, 3, do: Step 2a: Start with .j = f.(qj ) for j = 1, . . . , m and try to find real numbers c1,. . . ,cm such that fwa = c1.1 + · · · + cm.m. If such numbers are found, they become the m elements of row qi of transition matrix Wa, i.e., Wa(qi, qj) = cj for j = 1, . . . , m. Step 2b: If such numbers cannot be found, increment the number of states m = m + 1, and set .(qm) = wa, F(qm) = f(wa) (where F is the final distribution), and Wa(qi, qm) = 1. Step 3: Increment the index of the next unprocessed state i = i + 1. If i . m, go to Step 2. Step 4: The final step. Construct the initial distribution I by setting I(q1) = 1 and I(qj) = 0 for j = 2, 3, . . . , m. [Litow and Olivier 95] presents another inference algorithm that also yields a minimum state WFA and uses only additions and inner products. 704 7. Image Compression The real breakthrough in WFA compression came when a better inference algorithm was developed. This algorithm is the most important part of the WFA method. It generates a graph that may have more than the minimal number of states, but that has a small number of edges. The four transition matrices may be large, but they are sparse, resulting in better compression of the matrices and hence better compression of the image. The algorithm starts with a given finite-resolution image A, and generates its graph by matching subsquares of the image to the entire image or to other subsquares. An important point is that this algorithm can be lossy, with the amount of the loss controlled by a user-defined parameter G. Larger values of G “permit” the algorithm to match two parts of the image even if they poorly resemble each other. This naturally leads to better compression. The metric used by the algorithm to match two images (or image parts) f and g is the square of the L2 metric, a common measure where the distance dk(f, g) is defined as dk(f, g) = w [f(w) - g(w)]2 , where the sum is over all subsquares w. The algorithm tries to produce a small graph. If we denote the size of the graph by (size A), the algorithm tries to keep as small as possible the value of dk(f, fA) + G·(size A). The quantity m indicates the number of states, and there is a multiresolution image .i for each state i = 1, 2, . . . , m. The pseudocode of Figure 7.191 describes a recursive function make wfa(i, k, max) that tries to approximate .i at level k as well as possible by adding new edges and (possibly) new states to the graph. The function minimizes the value of the cost quantity cost = dk(.i, .i ) + G·s, where .i is the current approximation to .i and s is the increase in the size of the graph caused by adding new edges and states. If cost > max, the function returns ., otherwise, it returns cost. When the algorithm starts, m is set to 1 and . is set to f, where f is the function that needs to be approximated at level k. The algorithm then calls make wfa(1, k,.), which calls itself. For each of the four quadrants, the recursive call make wfa(i, k, max) tries to approximate (.i)a for a = 0, 1, 2, 3 in two different ways, as a linear combination of the functions of existing states (step 1), and by adding a new state and recursively calling itself (steps 2 and 3). The better of the two results is then selected (in steps 4 and 5). The algorithm constructs an initial distribution I = (1, 0, . . . , 0) and a final distribution Fi = .i() for all states i. WFA is a common acronym as, e.g., in “World Fellowship Activities.” 7.38 Finite Automata Methods 705 function make wfa(i, k, max); If max < 0, return .; cost ‹ 0; if k = 0, cost ‹ d0(f, 0) else do steps 1–5 with . = (.i)a for a = 0, 1, 2, 3; 1. Find r1, r2, . . . rm such that the value of cost1 ‹ dk-1(., r1.1 + · · · + rn.n) + G·s is small, where s denotes the increase in the size of the graph caused by adding edges from state i to states j with nonzero weights rj and label a, and dk-1 denotes the distance between two multiresolution images at level k - 1 (quite a mouthful). 2. Set m0 ‹ m, m ‹ m + 1, .m ‹ . and add an edge with label a and weight 1 from state i to the new state m. Let s denote the increase in the size of the graph caused by the new state and new edge. 3. Set cost2 ‹ G·s + make wfa(m, k - 1, min(max - cost, cost1) - G·s); 4. If cost2 . cost1, set cost ‹ cost + cost2; 5. If cost1 < cost2, set cost ‹ cost + cost1, remove all outgoing edges from states m0 + 1, . . . , m (added during the recursive call), as well as the edge from state i added in step 2. Set m ‹ m0 and add the edges from state i with label a to states j = 1, 2, . . . , m with weights rj whenever rj = 0. If cost . max, return(cost), else return(.). Figure 7.191: The WFA Recursive Inference Algorithm. Next, we discuss how the elements of the graph can be compressed. Step 2 creates an edge whenever a new state is added to the graph. They form a tree, and their weights are always 1, so each edge can be coded in four bits. Each of the four bits indicates which of two alternatives was selected for the label in steps 4–5. Step 1 creates edges that represent linear combinations (i.e., cases where subimages are expressed as linear combinations of other subimages). Both the weight and the endpoints need be stored. Experiments indicate that the weights are normally distributed, so they can be efficiently encoded with prefix codes. Storing the endpoints of the edges is equivalent to storing four sparse binary matrices, so run length encoding can be used. The WFA decoder reads vectors I and F and the four transition matrices Wa from the compressed stream and decompresses them. Its main task is to use them to reconstruct the original 2nΧ2n image A (precisely or approximately), i.e., to compute fA(w) for all strings w = a1a2 . . . an of size n. The original WFA decoding algorithm is fast but has storage requirements of order m4n, where m is the number of states of the graph. The algorithm consists of four steps as follows: Step 1: Set .p() = F(p) for all p . Q. Step 2: Repeat Step 3 for i = 1, 2, . . . , n. 706 7. Image Compression Step 3: For all p . Q, w = a1a2 . . . ai-1, and a = 0, 1, 2, 3, compute .p(aw) = q.Q Wa(p, q)·.q(w). Step 4: For each w = a1a2 . . . an compute fA(w) = q.Q I(q)·.q(w). This decoding algorithm was improved by Raghavendra Udupa, Vinayaka Pandit, and Ashok Rao [Udupa et al. 99]. They define a new multiresolution function .p for every state p of the WFA graph by .p() = I(p), .p(wa) = q.Q (Wa)q,p.q(w), for a = 0, 1, 2, 3, or, equivalently, .p(a1a2 . . . ak) = (I Wa1 . . . Wak)p, where .p(wa) is the sum of the weights of all the paths with label wa ending in p. The weight of a path starting at state i is the product of the initial distribution of i and the weights of the edges along the path wa. One result of this definition is that the intensity function f can be expressed as a linear combination of the new multiresolution functions .p, f(w) = q.Q .q(w)F(q), but there is a more important observation! Suppose that both .p(u) and .p(w) are known for all states p . Q of the graph and for all strings u and w. Then the intensity function of subsquare uw can be expressed as f(uw) = p.Q .p(u).p(w). This observation is used in the improved 6-step algorithm that follows to reduce both the space and time requirements. The input to the algorithm is a WFA A with resolution 2nΧ2n, and the output is an intensity function fA(w) for every string w up to length n. Step 1: Select nonnegative integers n1 and n2 satisfying n1 + n2 = n. Set .p = I(p) and .p = F(p) for all p . Q. Step 2: For i = 1, 2, . . . , n1 do Step 3. Step 3: For all p . Q and w = a1a2 . . . ai-1 and a = 0, 1, 2, 3, compute .p(aw) = q.Q (Wa)q,p.q(w). 7.38 Finite Automata Methods 707 Step 4: For i = 1, 2, . . . , n2 do Step 5. Step 5: For all p . Q and w = a1a2 . . . ai-1 and a = 0, 1, 2, 3, compute .p(aw) = q.Q (Wa)p,q.q(w). Step 6: For each u = a1a2 . . . an1 and w = b1b2 . . . bn2, compute fA(uw) = q.Q .q(u).q(w). The original WFA decoding algorithm is a special case of the above algorithm for n1 = 0 and n2 = n. For the case where n = 2l is an even number, it can be shown that the space requirement of this algorithm is of order m4n1 +m4n2. This expression has a minimum when n1 = n2 = n/2 = l, implying a space requirement of order m4l. Another improvement on the original WFA is the use of bintrees (Section 7.34.1). This idea is due to Ullrich Hafner [Hafner 95]. Recall that WFA uses quadtree methods to partition the image into a set of nonoverlapping subsquares. These can be called (in common with the notation used by IFS) the range images. Each range image is then matched, precisely or approximately, to a linear combination of other images that become the domain images. Using bintree methods to partition the image results in finer partitioning and an increase in the number of domain images available for matching. In addition, the compression of the transition matrices and of the initial and final distributions is improved. Each node in a bintree has two children compared to four children in a quadtree. A subimage can therefore be identified by a string of bits instead of a string of the digits 0–3. This also means that there are just two transition matrices W0 and W1. WFA compression works particularly well for color images, since it constructs a single WFA graph for the three color components of each pixel. Each component has its own initial state, but other states are shared. This normally saves many states because the color components of pixels in real images are not independent. Experience indicates that the WFA compression of a typical image may create, say, 300 states for the Y color component, 40 states for I, and only about 5 states for the Q component. Another property of WFA compression is that it is relatively slow for high-quality compression, since it builds a large automaton, but is faster for low-quality compression, since the automaton constructed in such a case is small. 7.38.2 Generalized Finite Automata The graphs constructed and used by WFA represent finite-state automata with “weighted inputs.” Without weights, finite automata specify multiresolution bi-level images. At resolution 2kΧ2k, a finite automaton specifies an image that is black in those subsquares whose addresses are accepted by the automaton. It is well known that every nondeterministic automaton can be converted to an equivalent deterministic one. This is not the case for WFA, where nondeterminism gives much more power. This is why an image with a smoothly varying gray scale, such as the one depicted in Figure 7.189, can be 708 7. Image Compression represented by a simple, two-state automaton. This is also why WFA can efficiently compress a large variety of grayscale and color images. For bi-level images the situation is different. Here nondeterminism might provide more concise description. However, experiments show that a nondeterministic automaton does not improve the compression of such images by much, and is slower to construct than a deterministic one. This is why the generalized finite automata (GFA) method was originally developed ([Culik and Valenta 96] and [Culik and Valenta 97a,b]). We also show how it is extended for color images. The algorithm to generate the GFA graph is completely different from the one used by WFA. In particular, it is not recursive (in contrast to the edge-optimizing WFA algorithm of Figure 7.191, which is recursive). GFA also uses three types of transformations, rotations, flips, and complementation, to match image parts. A GFA graph consists of states and of edges connecting them. Each state represents a subsquare of the image, with the first state representing the entire image. If quadrant q of the image represented by state i is identical (up to a scale factor) to the image represented by state j, then an edge is constructed from i to j and is labeled q. More generally, if quadrant q of state i can be made identical to state j by applying transformation t to the image of j, then an edge is constructed from i to j and is labeled q, t or (q, t). There are 16 transformations, shown in Figure 7.193. Transformation 0 is the identity, so it may be omitted when an edge is labeled. Transformations 1–3 are 90. rotations, and transformations 4–7 are rotations of a reflection of transformation 0. Transformations 8–15 are the reverse videos of 0–7, respectively. Each transformation t is thus specified by a 4-bit number. Figure 7.192: Image for Exercise 7.74. Figure 7.194 shows a simple bi-level image and its GFA graph. Quadrant 3 of state 0, for example, has to go through transformation 1 in order to make it identical to state 1, so there is an edge labeled (3, 1) from state 0 to state 1.  Exercise 7.74: Construct the GFA of the image of Figure 7.192 using transformations, and list all the resulting edges in a table. Images used in practice are more complex and less self-similar than the examples shown here, so the particular GFA algorithm discussed here (although not GFA in general) allows for a certain amount of data loss in matching subsquares of the image. The distance dk(f, g) between two images or subimages f and g of resolution 2kΧ2k is defined as dk(f, g) = w=a1a2...ak |f(w) - g(w)| 2kΧ2k . The numerator of this expression counts the number of pixels that differ in the two images, while the denominator is the total number of pixels in each image. The distance 7.38 Finite Automata Methods 709 0 1 2 3 0 1 2 3 0 1 2 3 0 1 2 3 01 23 01 23 01 23 01 23 0 1 2 3 0 1 2 3 0 1 2 3 0 1 2 3 01 23 01 23 01 23 01 23 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 Figure 7.193: Sixteen Image Transformations. 0,3 1,2 3,1 2,0 1,0 0,0 0,0 0,8 0,0 2,0 2,0 2,0 2,8 2,1 3,1 (0,0) (2,0) (2,1) (3,1) (a) 0 1 4 3 4 0 1 2 3 2 (0,0)(0,8) 0 1 2 3 3 3 3 2 2 2 1 1 1 0 0 0 0 2 (0,3) (1,2) (2,0) (3,1) (1,0) 3 (0,0)(2,0) (2,0)(2,8) (b) 1 Figure 7.194: A Five-State GFA. 710 7. Image Compression is thus the percentage of pixels that differ in the two images (we assume that 0 and 1 represent white and black pixels, respectively). The four-step algorithm described here for constructing the graph also uses an error parameter input by the user to control the amount of loss. Step 1: Generate state 0 as the entire image (i.e., subsquare ). We select the initial distribution as I = (1, 0, 0, . . . , 0), so state 0 will be the only one displayed. The final distribution is a vector of all ones. Step 2: Process each state as follows: Let the next unprocessed state be q, representing subsquare w. Partition w into its four quadrants w0, w1, w2, and w3, and perform Step 3 for each of the four wa’s. Step 3: Denote the image of subsquare wa by .. If . = 0, there is no edge from state q with label a. Otherwise, examine all the states generated so far and try to find a state p and a transformation t such that t(.p) (the image of state p transformed by t) will be similar enough to image .. This is expressed by dk(., t(.p)) . error. If such p and t are found, construct an edge q(a, t)p. Otherwise, add a new, unprocessed, state r to . and construct a new edge q(a, 0)r. Step 4: If there are any unprocessed states left, go to Step 2. Otherwise stop. Step 3 may involve many searches, since many states may have to be examined up to 16 times until a match is found. Therefore, the algorithm uses two versions for this search, a first fit and a best fit. Both approaches proceed from state to state, apply each of the 16 transformations to the state, and calculate the distance between each transformed state and .. The first fit approach stops when it finds the first transformed state that fits, while the best fit conducts the full search and selects the state and transformation that yield the best fit. The former is thus faster, while the latter produces better compression. The GFA graph for a real, complex image may have a huge number of states, so the GFA algorithm discussed here has an option where vector quantization is used. This option reduces the number of states (and thus speeds up the process of constructing the graph) by treating subsquares of size 8Χ8 differently. When this option is selected, the algorithm does not try to find a state similar to such a small subsquare, but rather encodes the subsquare with vector quantization (Section 7.19). An 8Χ8 subsquare has 64 pixels, and the algorithm uses a 256-entry codebook to code them. It should be noted that in general, GFA does not use quantization or any specific codebook. The vector quantization option is specific to the implementation discussed here. After constructing the GFA graph, its components are compressed and written on the compressed stream in three parts. The first part is the edge information created in step 3 of the algorithm. The second part is the state information (the indices of the states), and the third part is the 256 codewords used in the vector quantization of small subsquares. All three parts are arithmetically encoded. GFA decoding is done as in WFA, except that the decoder has to consider possible transformations when generating new image parts from existing ones. GFA has been extended to color images. The image is separated into individual bitplanes and a graph is constructed for each. However, common states in these graphs are stored only once, thereby improving compression. A little thinking shows that this works best for color images consisting of several areas with well-defined boundaries (i.e., discrete-tone or cartoon-like images). When such an image is separated into its bitplanes, 7.39 Iterated Function Systems 711 they tend to be similar. A subsquare q in bitplane 1, for example, may be identical to the same subsquare in, say, bitplane 3. The graphs for these bitplanes can thus share the state for subsquare q. The GFA algorithm is then applied to the bitplanes, one by one. When working on bitplane b, Step 3 of the algorithm searches through all the states of all the graphs constructed so far for the preceding b - 1 bitplanes. This process can be viewed in two different ways. The algorithm may construct n graphs, one for each bitplane, and share states between them. Alternatively, it may construct just one graph with n initial states. Experiments show that GFA works best for images with a small number of colors. Given an image with many colors, quantization is used to reduce the number of colors. 7.39 Iterated Function Systems Fractals have been popular since the 1970s and have many applications (see [Demko et al. 85], [Feder 88], [Mandelbrot 82], [Peitgen et al. 82], [Peitgen and Saupe 85], and [Reghbati 81] for examples). One such application, relatively underused, is data compression. Applying fractals to data compression is done by means of iterated function systems, or IFS. Two references to IFS are [Barnsley 88] and [Fisher 95]. IFS compression can be very efficient, achieving excellent compression factors (32 is not uncommon), but it is lossy and also computationally intensive. The IFS encoder partitions the image into parts called ranges; it then matches each range to some other part called a domain, and produces an affine transformation from the domain to the range. The transformations are written on the compressed stream, and they constitute the compressed image. We start with an introduction to two-dimensional affine transformations. 7.39.1 Affine Transformations In computer graphics, a complete two-dimensional image is built part by part and is normally edited before it is considered satisfactory. Editing is done by selecting a figure (part of the drawing) and applying a transformation to it. Typical transformations (Figure 7.195) are moving or sliding (translation), reflecting or flipping (mirror image), zooming (scaling), rotating, and shearing. The transformation can be applied to every pixel of the figure. Alternatively, it can be applied to a few key points that completely define the figure (such as the four corners of a rectangle), following which the figure is reconstructed from the transformed key points. We use the notation P = (x, y) for a two-dimensional point, and P* = (x*, y*) for the transformed point. The simplest linear transformation is x* = ax + cy, y* = bx+dy, in which each of the new coordinates is a linear combination of the two original coordinates. This transformation can be written P* = PT, where T is the 2Χ2 matrix a b c d. To understand the functions of the four matrix elements, we start by setting b = c = 0. The transformation becomes x* = ax, y* = dy. Such a transformation is called scaling. If applied to all the points of an object, all the x dimensions are scaled by a factor a, and all the y dimensions are scaled by a factor d. Note that a and d can also be 712 7. Image Compression Original Y scaled Y reflected X scaled Rotated Sheared X X* Clockwise rotation . . P P* Figure 7.195: Two-Dimensional Transformations. less than 1, causing shrinking of the object. If any of a or d equals -1, the transformation is a reflection. Any other negative values cause both scaling and reflection. Note that scaling an object by factors of a and d changes its area by a factor of aΧd, and that this factor is also the value of the determinant of the scaling matrix a 0 0 d. (Scaling, reflection, and other geometrical transformations can be extended to three dimensions, where they become much more complex.) Below are examples of matrices for scaling and reflection. In a, the y-coordinates are scaled by a factor of 2. In b, the x-coordinates are reflected. In c, the x dimensions are shrunk to 0.001 their original values. In d, the figure is shrunk to a vertical line: a = 1 0 0 2 , b= -1 0 0 1 , c= .001 0 0 1 , d= 0 0 0 1 . The next step is to set a = 1, d = 1 (no scaling or reflection), and explore the effect of b and c on the transformations. The transformation becomes x* = x+cy, y* = bx+y. We first select matrix 1 1 0 1 and use it to transform the rectangle whose four corners are (1, 0), (3, 0), (1, 1), and (3, 1). The corners are transformed to (1, 1), (3, 3), (1, 2), (3, 4). The original rectangle has been sheared vertically and transformed into a parallelogram. A similar effect occurs when we try the matrix 1 0 1 1. Thus, the quantities b and c are responsible for shearing. Figure 7.196 shows the connection between shearing and the operation of scissors. The word shearing comes from the concept of shear in mechanics. 7.39 Iterated Function Systems 713 Figure 7.196: Scissors and Shearing.  Exercise 7.75: Apply the shearing transformation 1-1 0 1to the four points (1, 0), (3, 0), (1, 1), and (3, 1). What are the transformed points? What geometrical figure do they represent? The next important transformation is rotation. Figure 7.195 illustrates rotation. It shows a point P rotated clockwise through an angle . to become P*. Simple trigonometry yields x = Rcos . and y = Rsin .. From this we get the expressions for x* and y*: x* = Rcos(. - .) = Rcos . cos . + Rsin . sin . = x cos . + y sin ., y* = Rsin(. - .) = -Rcos . sin . + Rsin . cos . = -x sin . + y cos .. Thus the rotation matrix in two dimensions is cos . -sin . sin . cos . , (7.63) which also equals cos . 0 0 cos . 1 -tan . tan . 1 . This proves that any rotation in two dimensions is a combination of scaling (and, perhaps, reflection) and shearing, an unexpected result that’s true for all angles satisfying tan . = .. Matrix T1 below rotates anticlockwise. Matrix T2 reflects about the line y = x, and matrix T3 reflects about the line y = -x. Note the determinants of these matrices. In general, a determinant of +1 indicates pure rotation, whereas a determinant of -1 indicates pure reflection. (As a reminder, det a b c d = ad - bc.) T1 = cos . sin . -sin . cos . , T2 = 0 1 1 0 , T3 = 0 -1 -1 0 . A 90. Rotation In the case of a 90. clockwise rotation, the rotation matrix is cos(90) -sin(90) sin(90) cos(90) = 0 -1 1 0 . (7.64) 714 7. Image Compression A point P = (x, y) is thus transformed to the point (y,-x). For a counterclockwise 90. rotation, (x, y) is transformed to (-y, x). This is called the negate and exchange rule. Translations Unfortunately, our simple 2 Χ 2 matrix cannot generate all the necessary transformations! Specifically, it cannot generate translation. This is proved by realizing that any object containing the origin will, after any of the transformations above, still contain the origin (the result of (0, 0) Χ T is (0, 0) for any matrix T). One way to implement translation (which can be expressed by x* = x + m, y* = y+n), is to generalize our transformations to P* = PT+(m, n), where T is the familiar 2 Χ 2 transformation matrix a b c d. A more elegant approach, however, is to stay with P* = PT and to generalize T to the 3Χ3 matrix T =.. a b 0 c d 0 m n 1.. . This approach is called homogeneous coordinates and is commonly used in projective geometry. It makes it possible to unify all the two-dimensional transformations within one matrix. Notice that only six of the nine elements of matrix T are variables. Our points should now be the triplets P = (x, y, 1). It is easy to see that the transformations discussed above can change lengths and angles. Scaling changes the lengths of objects. Rotation and shearing change angles. One property that is preserved, though, is parallel lines. A pair of parallel lines will remain parallel after any scaling, reflection, rotation, shearing, and translation. A transformation that preserves parallelism is called affine. The final conclusion of this section is that any affine two-dimensional transformation can be fully specified by only six numbers! Affine transformations can be defined in different ways. One important definition is that a transformation of points in space is affine if it preserves barycentric sums of the points. A barycentric sum of points Pi has the form wiPi, where wi are numbers and wi = 1, so if P = wiPi and if wi = 1, then any affine transformation T satisfies TP = n 1 wiTPi. 7.39.2 IFS Definition A simple example of IFS is the set of three transformations T1 =.. .5 0 0 0 .5 0 8 8 1 .. , T2 =.. .5 0 0 0 .5 0 96 16 1.. , T3 =.. .5 0 0 0 .5 0 120 60 1.. . (7.65) We first discuss the concept of the fixed point. Imagine the sequence P1 = P0T1, P2 = P1T1, . . ., where transformation T1 is applied repeatedly to create a sequence of 7.39 Iterated Function Systems 715 points P1, P2, . . .. We start with an arbitrary point P0 = (x0, y0) and examine the limit of the sequence P1 = P0T1, P2 = P1T1,. . . . It is easy to show that the limit of Pi for large i is the fixed point (2m, 2n) = (16, 16) (where m and n are the translation factors of matrix T1, Equation (7.65)) regardles of the values of x0 and y0. Point (16, 16) is called the fixed point of T1, and it does not depend on the particular starting point P0 selected. Proof: P1 = P0T1 = (.5x0+8, .5y0+8), P2 = P1T1 = (0.5(0.5x0+8)+8, .5(0.5y0+ 8)+8). It is easy to see (and to prove by induction) that xi = 0.5ix0+0.5i-18+0.5i-28+ · · ·+0.518+8. In the limit xi = 0.5ix0+8.j =0 0.5j = 0.5ix0+8Χ2, which approaches the limit 8 Χ 2 = 16 for large i regardless of x0. Now it is easy to show that for the transformations above, with scale factors of 0.5 and no shearing, each new point in the sequence moves half the remaining distance toward the fixed point. Given a point Pi = (xi, yi), the point midway between Pi and the fixed point (16, 16) is xi + 16 2 , yi + 16 2 = (0.5xi + 8, 0.5yi + 8) = (xi+1, yi+1) = Pi+1. Consequently, for the particular transformations above there is no need to use the transformation matrix. At each step of the iteration, point Pi+1 is obtained by (Pi + (2m, 2n))/2. For other transformations, matrix multiplication is necessary to compute point Pi+1. In general, every affine transformation where the scale and shear factors are less than 1 has a fixed point, but it may not be easy to find it. The principle of IFS is now easy to describe. A set of transformations (an IFS code) is selected. A sequence of points is computed and plotted by starting with an arbitrary point P0, selecting a transformation from the set at random, and applying it to P0, transforming it into a point P1, and then randomly selecting another transformation and applying it to P1, thereby generating point P2, and so on. Every point is plotted on the screen as it is calculated, and gradually, the object begins to take shape before the viewer’s eyes. The shape of the object is called the IFS attractor, and it depends on the IFS code (the transformations) selected. The shape also depends slightly on the particular selection of P0. It is best to choose P0 as one of the fixed points of the IFS code (if they are known in advance). In such a case, all the points in the sequence will lie inside the attractor. For any other choice of P0, a finite number of points will lie outside the attractor, but eventually they will move into the attractor and stay there. It is surprising that the attractor does not depend on the precise order of the transformations used. This result has been proved by the mathematician John Elton. Another surprising property of IFS is that the random numbers used don’t have to be uniformly distributed; they can be weighted. Transformation T1, for example, may be selected at random 50% of the time, transformation T2, 30%, and T3, 20%. The shape being generated does not depend on the probabilities, but the computation time does. The weights should add up to 1 (a normal requirement for a set of mathematical weights), and none can be zero. 716 7. Image Compression The three transformations of Equation (7.65) above create an attractor in the form of a Sierpi΄nski triangle (Figure 7.197a). The translation factors determine the coordinates of the three triangle corners. The six transformations of Table 7.198 create an attractor in the form of a fern (Figure 7.197b). The notation used in Table 7.198 is a b c d+ m n. The Sierpi΄nski triangle, also known as the Sierpi΄nski gasket (Figure 7.197a), is defined recursively. Start with any triangle, find the midpoint of each edge, connect the three midpoints to obtain a new triangle fully contained in the original one, and cut the new triangle out. The newly created hole now divides the original triangle into three smaller ones. Repeat the process on each of the smaller triangles. At the limit, there is no area left in the triangle. It resembles Swiss cheese without any cheese, just holes. (a) (b) Figure 7.197: (a) Sierpi΄nski Triangle. (b) A Leaf. The program of Figure 7.199 calculates and displays IFS attractors for any given set of transformations. It runs on the Macintosh computer (because of changes in the Macintosh operating system, this program no longer runs and should be considered pseudocode). 7.39.3 IFS Principles Before we describe how IFS is used to compress real-life images, let’s look at IFS from a different point of view. Figure 7.200 shows three images: a person, the letter “T”, and the Sierpi΄nski gasket (or triangle). The first two images are transformed in a special way. Each image is shrunk to half its size, then copied three times, and the three copies are arranged in the shape of a triangle. When this transformation is applied a few times to an image, it is still possible to discern the individual copies of the original image. However, when it is applied many times, the result is the Sierpi΄nski gasket (or something very close to it, depending on the number of iterations and on the resolution of the output device). The point is that each transformation shrinks the image (the transformations 7.39 Iterated Function Systems 717 a b c d m n a b c d m n 1: 0 -28 0 29 151 92 4: 64 0 0 64 82 6 2: -2 37 -31 29 85 103 5: 0 -80 -22 1 243 151 3: -1 18 -18 -1 88 147 6: 2 -48 0 50 160 80 Table 7.198: All numbers are shown as integers, but a, b, c, and d should be divided by 100, to make them less than 1. The values m and n are the translation factors. PROGRAM IFS; USES ScreenIO, Graphics, MathLib; CONST LB = 5; Width = 490; Height = 285; (* LB=left bottom corner of window *) VAR i,k,x0,y0,x1,y1,NumTransf: INTEGER; Transf: ARRAY[1..6,1..10] OF INTEGER; Params:TEXT; filename:STRING; BEGIN (* main *) Write(’params file=’); Readln(filename); Assign(Params,filename); Reset(Params); Readln(Params,NumTransf); FOR i:=1 TO NumTransf DO Readln(Params,Transf[1,i],Transf[2,i],Transf[3,i], Transf[4,i],Transf[5,i],Transf[6,i]); OpenGraphicWindow(LB,LB,Width,Height,’IFS shape’); SetMode(paint); x0:=100; y0:=100; REPEAT k:=RandomInt(1,NumTransf+1); x1:=Round((x0*Transf[1,k]+y0*Transf[2,k])/100)+Transf[5,k]; y1:=Round((x0*Transf[3,k]+y0*Transf[4,k])/100)+Transf[6,k]; Dot(x1,y1); x0:=x1; y0:=y1; UNTIL Button()=TRUE; ScBOL; ScWriteStr(’Hit a key & close this window to quit’); ScFreeze; END. Figure 7.199: Calculate and Display IFS Attractors. 718 7. Image Compression are contractive), so the final result does not depend on the shape of the original image. The shape can be that of a person, a letter, or anything else; the final result depends only on the particular transformation applied to the image. A different transformation will create a different result, which again will not depend on the particular image being transformed. Figure 7.200d, for example, shows the results of transforming the letter “T” by reducing it, making three copies, arranging them in a triangle, and flipping the top copy. The final image obtained at the limit, after applying a certain transformation infinitely many times, is called the attractor of the transformation. (An attractor is a state to which a dynamical system evolves after a long enough time.) The following sets of numbers create especially interesting patterns. 1. A frond. 5 0 -28 0 29 151 92 64 0 0 64 82 6 -2 37 -31 29 85 103 17 -51 -22 3 183 148 -1 18 -18 -1 88 147 2. A coastline 4 -17 -26 34 -12 84 53 25 -20 29 17 192 57 35 0 0 35 49 3 25 -6 6 25 128 28 3. A leaf (Figure 7.197b) 4 2 -7 -2 48 141 83 40 0 -4 65 88 10 -2 45 -37 10 82 132 -11 -60 -34 22 237 125 4. A Sierpi΄nski Triangle 3 50 0 0 50 0 0 50 0 0 50 127 79 50 0 0 50 127 0  Exercise 7.76: The three affine transformations of example 4 above (the Sierpi΄nski triangle) are different from those of Equation (7.65). What is the explanation? The result of each transformation is an image containing all the images of all the previous transformations. If we apply the same transformation many times, it is possible to zoom on the result, to magnify it many times, and still see details of the original images. In principle, if we apply the transformation an infinite number of times, the final result will show details at any magnification. It will be a fractal. The case of Figure 7.200c is especially interesting. It seems that the original image is simply shown four times, without any transformations. A little thinking, however, shows that our particular transformation transforms this image to itself. The original image is already the Sierpi΄nski gasket, and it gets transformed to itself because it is self-similar.  Exercise 7.77: Explain the geometrical meaning of the combined three affine transformations below and show the attractor they converge to w1 x y = 1/2 0 0 1/2 x y , w2 x y = 1/2 0 0 1/2 x y + 0 1/2 , w3 x y = 1/2 0 0 1/2 x y + 1/2 0 . 7.39 Iterated Function Systems 719 T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T T (d) (c) (b) (a) Figure 7.200: Creating the Sierpi΄nski Gasket. Figure 7.201: A Self-Similar Image. 720 7. Image Compression The Sierpi΄nski gasket is therefore easy to compress because it is self-similar; it is easy to find parts of it that are identical to the entire image. In fact, every part of it is identical to the entire image. Figure 7.201 shows the bottom-right part of the gasket surrounded by dashed lines. It is easy to see the relation between this part and the entire image. Their shapes are identical, up to a scale factor. The size of this part is half the size of the image, and we know where it is positioned relative to the entire image (we can measure the displacement of its bottom-left corner from the bottom-left corner of the entire image). This points to a possible way to compress real images. If we can divide an image into parts such that each part is identical (or at least very close) to the entire image up to a scale factor, then we can highly compress the image by IFS. All that we need is the scale factor (actually two scale factors, in the x and y directions) and the displacement of each part relative to the entire image [the (x, y) distances between a corner of the part and the same corner of the image]. Sometimes we may find a part of the image that has to be reflected in order to become identical to the entire image. In such a case we also need the reflection coefficients. Thus, we can compress an image by figuring out the transformations that transform each part (called “range”) into the entire image. The transformation for each part is expressed by a few numbers, and these numbers become the compressed stream. It is easy to see that this simple approach will not work for real-life images. Such images are complex, and it is generally impossible to divide such an image into parts that will all be identical (or even very close) to the entire image. A different approach is needed to make IFS practical. The approach used by any practical IFS algorithm is to partition the image into nonoverlapping parts called ranges. They can be of any size and shape, but in practice it is easiest to work with squares, rectangles, or triangles. For each range Ri, the encoder has to find a domain Di that’s very similar, or even identical in shape, to the range but is bigger. Once such a domain is found, it is easy to figure out the transformation wi that will transform the domain into the range Ri = wi(Di). Two scale factors have to be determined (the scaling is shrinking, since the domain is bigger than the range) as well as the displacement of the domain relative to the range [the (x, y) distances between one corner of the domain and the same corner of the range]. Sometimes, the domain has to be rotated and/or reflected to make it identical to the range, and the transformation should, of course, include these factors as well. This approach to IFS image compression is called PIFS (for partitioned IFS). 7.39.4 IFS Decoding Before looking into the details of PIFS encoding, let’s try to understand how the PIFS decoder works. All that the decoder has is the set of transformations, one per range. It does not know the shapes of any ranges or domains. In spite of this, decoding is very simple. It is based on the fact, mentioned earlier, that a contractive transformation creates a result that does not depend on the shape of the initial image used. We can therefore create any range Ri by applying contractive transformation wi many times to any bigger shape Di (except an all-white shape). The decoder therefore starts by setting all the domains to arbitrary shapes (e.g., it can initially set the entire image to black). It then goes into a loop where in each iteration it applies every transformation wi once. The first iteration applies the transformations 7.39 Iterated Function Systems 721 to domains Di that are all black. This creates ranges Ri that may already, after this single iteration, slightly resemble the original ranges. This iteration changes the image from the initial all black to something resembling the original image. In the second iteration the decoder applies again all the wi transformations, but this time they are applied to domains that are no longer all black. The domains already somewhat resemble the original ones, so the second iteration results in better-looking ranges, and thus in a better image. Experience shows that only 8–10 iterations are normally needed to get a result that closely resembles the original image. It is important to realize that this decoding process is resolution independent! Normally, the decoder starts with an initial image whose size is identical to that of the original image. It can, however, start with an all-black image of any size. The affine transformations used to encode the original image do not depend on the resolution of the image or on the resolutions of the ranges. Decoding an image at, say, twice its original size will create a large, smooth image, with details not seen in the original and without pixelization (without jagged edges or “fat” pixels). The extra details will, of course, be artificial. They may not be what one would see when looking at the original image through a magnifying glass, but the point is that PIFS decoding is resolution independent; it creates a natural-looking image at any size, and it does not involve pixelization. The resolution-independent nature of PIFS decoding also means that we have to be careful when measuring the compression performance. After compressing a 64-Kb image into, say, 2 Kb, the compression factor is 32. Decoding the 2-Kb compressed file into a large, 2 Mb, image (with a lot of artificial detail) does not mean that we have changed the compression factor to 2M/2K=1024. The compression factor is still the same 32. 7.39.5 IFS Encoding PIFS decoding is therefore easy, if somewhat magical, but we still need to see the details of PIFS encoding. The first point to consider is how to select the ranges and find the domains. The following is a straightforward way of doing this. Suppose that the original image has resolution 512Χ512. We can select as ranges the nonoverlapping groups of 16Χ16 pixels. There are 32Χ32 = 1024 such groups. The domains should be bigger than the ranges, so we may select as domains all the 32Χ32 groups of pixels in the image (they may, of course, overlap). There are (512-31)Χ(512- 31) = 231361 such groups. The encoder should compare each range with all 231,361 domains. Each comparison involves eight steps, because a range may be identical to a rotation or a reflection of a domain (Figure 7.193). The total number of steps in comparing ranges and domains is therefore 1024Χ231361Χ8 = 1,895,309,312. If each step takes a microsecond, the total time required is 1,895 seconds, or about 31 minutes.  Exercise 7.78: Repeat the computation above for a 256 Χ 256 image with ranges of size 8 Χ 8 and domains of size 16 Χ 16. If the encoder is looking for a domain to match range Ri and it is lucky to find one that’s identical to Ri, it can proceed to the next range. In practice, however, domains identical to a given range are very rare, so the encoder has to compare all 231,361Χ8 domains to each range Ri and select the one that’s closest to Ri (PIFS is, in general, a lossy compression method). We therefore have to answer two questions: When is 722 7. Image Compression a domain identical to a range (remember: they have different sizes) and how do we measure the “distance” between a domain and a range? To compare a 32Χ32-pixel domain to a 16Χ16-pixel range, we can either choose one pixel from each 2Χ2 square of pixels in the domain (this is called subsampling) or average each 2Χ2 square of pixels in the domain and compare it to one pixel of the range (averaging). To decide how close a range Ri is to a domain Dj , we have to use one of several metrics. A metric is a function that measures “distance” between, or “closeness” of, two mathematical quantities. Experience recommends the use of the rms (root mean square) metric Mrms(Ri,Dj) =  x,y Ri(x, y) - Dj(x, y)2 . (7.66) This involves a square root calculation, so a simpler metric may be Mmax(Ri,Dj) = max |Ri(x, y) - Dj(x, y)| (the largest difference between a pixel of Ri and a pixel of Dj ). Whatever metric is used, a comparison of a range and a domain involves subsampling (or averaging) followed by a metric calculation. After comparing range Ri to all the (rotated and reflected) domains, the encoder selects the domain with the smallest metric and determines the transformation that will bring the domain to the range. This process is repeated for all ranges. Even this simple way of matching parts of the image produces excellent results. Compression factors are typically in the 15–32 range and data loss is minimal.  Exercise 7.79: What is a reasonable way to estimate the amount of image information lost in PIFS compression? The main disadvantage of this method of determining ranges and domains is the fixed size of the ranges. A method where ranges can have different sizes may lead to better compression and less loss. Imagine an image of a hand with a ring on one finger. If the ring happens to be inside a large range Ri, it may be impossible to find any domain that will even come close to Ri. Too much data may be lost in such a case. On the other hand, if part of an image is fairly uniform, it may become a large range, since there is a better chance that it will match some domain. Clearly, large ranges are preferable, since the compressed stream contains one transformation per range. Therefore, quadtrees offer a good solution. Quadtrees: We start with a few large ranges, each a subquadrant. If a range does not match well any domain (the metric between it and any domain is greater than a user-controlled tolerance parameter), it is divided into four subranges, and each is matched separately. As an example, consider a 256Χ256-pixel image. We can choose for domains all the image squares of size 8, 12, 16, 24, 32, 48, and 64 pixels. We start with ranges that are subquadrants of size 32 Χ 32. Each range is compared with domains that are larger than itself (48 or 64) pixels. If a range does not match well, it is divided into four quadrants of size 16 Χ 16 each, and each is compared, as a new range, with all domains of sizes 24, 32, 48, and 64 pixels. This process continues until all ranges 7.39 Iterated Function Systems 723 have been matched to domains. Large ranges result in better compression, but small ranges are easier to match, because they contain few adjacent pixels, and we know from experience that adjacent pixels tend to be highly correlated in real-life images. 7.39.6 IFS for Grayscale Images Up until now we have assumed that our transformations have to be affine. The truth is that any contractive transformations, even nonlinear ones, can be used for IFS. Affine transformations are used simply because they are linear and therefore computationally simple. Also, up to now we have assumed a monochromatic image, where the only problem is to determine which pixels should be black. IFS can easily be extended to compress grayscale images (and therefore also color images; see below). The problem here is to determine which pixels to paint, and what gray level to paint each. Matching a domain to a range now involves the intensities of the pixels in both. Whatever metric is employed, it should use only those intensities to determine the “closeness” of the domain and the range. Assume that a certain domain D contains n pixels with gray levels a1 . . . an, and the IFS encoder tries to match D to a range R containing n pixels with gray levels b1, . . . , bn. The rms metric, mentioned earlier, works by finding two numbers, r and g (called the contrast and brightness controls), that will minimize the expression Q = n 1 (r · ai + g) - bi2 . (7.67) This is done by solving the two equations .Q/.r = 0 and .Q/.g = 0 for the unknowns r and g (see details below). Minimizing Q minimizes the difference in contrast and brightness between the domain and the range. The value of the rms metric is .Q [compare with Equation (7.66)]. When the IFS encoder finally decides which domain to associate with the current range, it has to figure out the transformation w between them. The point is that r and g should be included in the transformation, so that the decoder will know what gray level to paint the pixels when the domain is recreated in successive decoding iterations. It is common to use transformations of the form w.. x y z.. =.. a b 0 c d 0 0 0 r.. .. x y z.. +.. l m g.. . (7.68) A pixel (x, y) in the domain D is now given a third coordinate z (its gray level) and is transformed into a pixel (x*, y*, z*) in the range R, where z* = z · r + g. Transformation (7.68) has another property. It is contractive if r < 1, regardless of the scale factors. Any compression method for grayscale images can be extended to color images. It is only necessary to separate the image into three color components (preferably YIQ) and compress each individually as a grayscale image. This is how IFS can be applied to the compression of color images. The next point that merits consideration is how to write the coefficients of a transformation w on the compressed stream. There are three groups of coefficients, the scale factors a and d, the reflection/rotation factors a, b, c, and d, the displacements l and m, 724 7. Image Compression and the contrast/brightness controls r and g. If a domain is twice as large as a range, then the scale factors are always 0.5 and consequently do not have to be written on the compressed stream. If the domains and ranges can have several sizes, then only certain scale factors are possible, and they can be encoded either arithmetically or with some prefix code. The particular rotation or reflection of the domain relative to the range can be coded with three bits, since there are only eight rotations/reflections possible. The displacement can be encoded by encoding the positions and sizes of the domain and range. The quantities r and g are not distributed in any uniform way, and they are also real (floating-point) numbers that can have many different values. They should thus be quantized, i.e., converted to an integer in a certain range. Experience shows that the contrast r can be quantized into a 4-bit or 5-bit integer (i.e., 16 or 32 contrast values are enough in practice), whereas the brightness g should become a 6-bit or 7-bit integer (resulting in 64 or 128 brightness values). Here are the details of computing r and g that minimize Q and then calculating the (minimized) Q and the rms metric. From Equation (7.67), we get .Q .g = 0 › 2(r · ai + g - bi) = 0 › ng + (r · ai - bi) = 0, g = 1 n  bi - r ai, (7.69) and .Q .r = 0 › 2(r · ai + g - bi)ai = 0 › (r · a2i + g · ai - aibi) = 0, r a2i + 1 n  bi - r ai ai - aibi = 0, r  a2i - 1 n  ai2= aibi - 1 n ai bi, r = aibi - 1 n aibi a2i - 1 n (ai)2 = naibi -aibi na2i - (ai)2 . (7.70) From the same Equation (7.67), we also get the minimized Q Q = n 1 (r · ai + g - bi)2 = (r2a2i + g2 + b2i + 2rgai - 2raibi - 2gbi) = r2 a2i + ng2 + b2i + 2rg ai - 2r aibi - 2g bi. (7.71) The following steps are needed to calculate the rms metric: 1. Compute the sums ai and a2i for all domains. 2. Compute the sums bi and b2i for all ranges. 3. Every time a range R and a domain D are compared, compute: 3.1. The sum aibi. 7.40 Spatial Prediction 725 3.2. The quantities r and g from Equations (7.69) and (7.70) using the five sums above. Quantize r and g. 3.3. Compute Q from Equation (7.71) using the quantized r, g and the five sums. The value of the rms metric for these particular R and D is .Q. Finally, Figures 7.202 and 7.203 are pseudocode algorithms outlining two approaches to IFS encoding. The former is more intuitive. For each range R it selects the domain that’s closest to R. The latter tries to reduce data loss by sacrificing compression ratio. This is done by letting the user specify the minimum number T of transformations to be generated. (Each transformation w is written on the compressed stream using roughly the same number of bits, so the size of that stream is proportional to the number of transformations.) If every range has been matched, and the number of transformations is still less than T, the algorithm continues by taking ranges that have already been matched and partitioning them into smaller ones. This increases the number of transformations but reduces the data loss, since smaller ranges are easier to match with domains. He gathered himself together and then banged his fist on the table. “To hell with art, I say.” “You not only say it, but you say it with tiresome iteration,” said Clutton severely. —W. Somerset Maugham, Of Human Bondage 7.40 Spatial Prediction Image compression by spatial prediction is a combination of JPEG and various fractalbased image compression algorithms. The method is described in [Feig et al. 95]. The principle is to select an integer N, partition the input image into blocks of N ΧN pixels each, and compress each block with JPEG or with a simple fractal method depending on which is better. Compressing the block with JPEG results in a zigzag sequence of N Χ N transform coefficients which are quantized and further compressed as described in Section 7.10.4. In contrast, the fractal compression of a block results in five integers as shown later in this section. The method described in Section 7.39 is perhaps the simplest fractal-based approach to image compression. It is based on the observation that any part of an image may look very similar to other parts, or to transformed versions of other parts. Thus, a given part A of an image may be very close to a scaled, rotated, and reflected version of part B. The method partitions the image into blocks and tries to match the current block A to a transformed version of a previously-encoded block. If (a transformed version of) a block B provides a sufficiently-close match, the location of B, as well as the transform parameters, are written on the output and become the compressed data of the current block. 726 7. Image Compression t:=some default value; [t is the tolerance] push(entire image); [stack contains ranges to be matched] repeat R:=pop(); match all domains to R, find the one (D) that’s closest to R, pop(R); if metric(R,D) 0 and .2 is finite. Note how the function in Figure 8.2c, obtained by adding the simple functions in Figure 8.2a,b, starts resembling a square pulse. It turns out that we can bring it closer to a square pulse (like the one in Figure 8.1a) by adding (1/5) sin(2.(5f)t), (1/7) sin(2.(7f)t), and so on. We say that the Fourier series of a square wave with amplitude A and frequency f is the infinite sum A . k=1,3,5,... 1 k sin(2.kft), where successive terms have smaller and smaller amplitudes. Here are two examples that show the relation between a function g(t) and its Fourier expansion G(f). The first example is the step function of Figure 8.4a, defined by g(t) = ./2, for 2k. . t < (2k + 1)., -./2, for (2k + 1). . t < (2k + 2)., where k is any integer (positive or negative). The Fourier expansion of g(t) is G(f) = 2 . k=0 sin[f(2k + 1)] 2k + 1 = 2 sin f + 2 sin 3f 3 + 2 sin 5f 5 + · · · . Figure 8.5a shows the first three terms of this series, and Figure 8.5b shows three partial sums, of the first four, eight, and 13 terms. It is obvious that these partial sums quickly approach the original function. The second example is the sawtooth function of Figure 8.4b, defined by g(t) = t/2 for every interval [2k., (2k + 1).). Its Fourier expansion is G(f) = . 2 - . k=1 sin(k f) k . Figure 8.6 shows four partial sums, of the first three, five, nine, and 17 terms, of this expansion. 8.3 The Uncertainty Principle 737 (a) (b) t 2. t . 2. ./2 -./2 -. ./2 . Figure 8.4: Two Functions g(t) to be Transformed. 8.3 The Uncertainty Principle These examples illustrate an important relation between the time and frequency domains; namely, they are complementary. Each of them complements the other in the sense that when one of them is localized the other one must be global. Consider, for example, a pure sine wave. It has one frequency, so it is well localized in the frequency domain. However, it is infinitely long, so in the time domain it is global. On the other hand, a function may be localized in the time domain, such as the single spike of Figure 8.38a, but it will invariably be global in the frequency domain; its Fourier expansion will contain many (possibly even infinitely many) frequencies. This relation between the time and frequency domains makes them complementary, and it is popularly described by the term uncertainty relation or uncertainty principle. The Heisenberg uncertainty principle, first recognized in 1927 by Werner Heisenberg, is a very important physical principle. It says that position and momentum are complementary. The better we know the position of a particle of matter, the less certain we are of its momentum. The reason for this relation is the way we measure positions. In order to locate a particle in space, we have to see it (with our eyes or with an instrument). This requires shining light on the particle (either visible light or some other wavelength), and it is this light that disturbs the particle, moves it from its position and thus changes its momentum. We think of light as consisting of small units, photons, that don’t have any mass but have momentum. When a photon hits a particle, it gives the particle a “kick,” which moves it away. The larger the particle, the smaller the effects of the kick, but in principle, the complementary relation between position and momentum exists even for macroscopic objects. It is important to realize that the uncertainty principle is part of nature; it is not just a result of our imperfect knowledge or primitive instruments. It is expressed by the relation .x.p . ―h 2 = h 4. , where .x and .p are the uncertainties in the position and momentum of the particle, 738 8. Wavelet Methods 1 -. . 2. -1 -2. (a) (b) -./2 ./2 -2. -. . 2. Figure 8.5: The First Three Terms and Three Partial Sums. 8.3 The Uncertainty Principle 739 -2. 2. ./2 . Figure 8.6: Four Partial Sums. and h is the Planck constant (6.626176Χ10-27erg·sec). The point is that h is very small, so ―h def = h 2. = 1.05Χ10-34 = 0.000000000000000000000000000000000105 joule·sec. (An erg and a joule are units of energy whose precise values are irrelevant for this discussion.) This is why we don’t notice the position/momentum uncertainty in everyday life. A simple example should make this clear. Suppose that the mass of a moving ball is 0.15 Kg and we measure its velocity to a precision of 0.05 m/sec (that is about 1 mph). Momentum is the product of mass and velocity, so if the uncertainty in the velocity of the ball is 0.05 m/sec, the uncertainty in its momentum is .p = 0.15Χ0.05 = 0.0075Kg·m/sec The uncertainty principle says that .xΧ0.0075 . 1.05Χ10-34/2, which implies that the uncertainty in the position of the ball is on the order of 70 Χ 10-34 m or 7 Χ 10-30 mm, too small to measure! 740 8. Wavelet Methods The uncertainty principle is a special case of the principle of complementarity. Position and momentum are complementary quantities, but there are other pairs of physical quantities that are complementary, such as time and energy. While we are discussing physics, here is another similarity between the time and frequency domains and physical reality. Real objects (i.e., objects with mass) have a property called the “wave particle duality.” This is true for all objects but is noticeable and can be measured only for microscopic particles. This property means that every object has a wave associated with it. The amplitude of this wave (actually, the square of the absolute value of the amplitude) in a given region describes the probability of finding the object in that region. When we observe the object, it will normally be found in the region where its wave function has the highest amplitude. The wave associated with an object is generally not spread throughout the entire universe, but is confined to a small region in space. This “localized” wave is called a wave packet. Such a packet does not have one particular wavelength, but consists of waves of many wavelengths. The momentum of an object depends on its wavelength, so an object that is localized in space doesn’t have a specific, well-defined momentum; it has many momenta. In other words, there’s an uncertainty in the momentum of the particle. Nuggets of Uncertainty A full discussion of Heisenberg’s uncertainty principle may be found in the Appendix. Then again, it may not. Uncertainty is the very essence of romance. It’s what you don’t know that intrigues you. 8.4 Fourier Image Compression We now apply the concepts of time and frequency domains to digital images. Imagine a grayscale photograph scanned line by line. For all practical purposes we can assume that the photograph has infinite resolution (its shades of gray can vary continuously). An ideal scan would therefore result in an infinite sequence of numbers that can be considered the values of a (continuous) intensity function I(t). In practice, we can store only a finite amount of data in memory, so we have to select a finite number of values I(1) through I(n). This process is known as sampling. Intuitively, sampling seems a trade-off between quality and price. The bigger the sample, the better the quality of the final image, but more hardware (more memory and higher screen resolution) is required, resulting in higher costs. This intuitive conclusion, however, is not entirely true. Sampling theory tells us that we can sample an image and reconstruct it later in memory without loss of quality if we can do the following: 1. Transform the intensity function from the time domain I(t) to the frequency domain G(f). 2. Find the maximum frequency fm. 8.4 Fourier Image Compression 741 3. Sample I(t) at a rate slightly higher than 2fm (e.g., if fm = 22,000 Hz, generate samples at the rate of 44,100 Hz). 4. Store the sampled values in the bitmap. The resulting image would be equal in quality to the original one on the photograph. There are two points to consider. The first is that fm could be infinite. In such a case, a value fm should be selected such that frequencies that are greater than fm do not contribute much (have low amplitudes). There is some loss of image quality in such a case. The second point is that the bitmap (and consequently, the resolution) may be too small for the sample generated in step 3. In such a case, a smaller sample has to be taken, again resulting in a loss of image quality. The result above was proved by Harry Nyquist [Nyquist 28], and the quantity 2fm is called the Nyquist rate. It is used in many practical situations. The range of human hearing, for instance, is between 16 Hz and 22,000 Hz. When sound is digitized at high quality (such as music recorded on a CD), it is sampled at the rate of 44,100 Hz. Anything lower than that results in distortions. Section 7.1 discusses the application of this theorem to images (two-dimensional signals). Why is it that sampling at the Nyquist rate is enough to restore the original signal? It seems that sampling ignores the behavior of the analog signal between samples, and can therefore miss important information. What guarantees that the signal will not go up or down considerably between consecutive samples? In principle, such behavior may happen, but in practice, all analog signals have a frequency response limit because they are created by sources (such as microphone, speaker, or mouth) whose response speed is limited. Thus, the rate at which a real signal can change is limited, thereby making it possible to predict the way it will vary from sample to sample. We say that the finite bandwidth of real signals is what makes their digitization possible. Fourier is a mathematical poem. —William Thomson (Lord Kelvin) The Fourier transform is useful and popular, having applications in many areas. It has, however, one drawback: It shows the frequency content of a function f(t), but it does not specify where (i.e., for what values of t) the function has low or high frequencies. The reason for this is that the basis functions (sine and cosine) used by this transform are infinitely long. They pick up the different frequencies of f(t) regardless of where they are located. A better transform should specify the frequency content of f(t) as a function of t. Instead of producing a one-dimensional function (in the continuous case) or a onedimensional array of numbers (in the discrete case), it should produce a two-dimensional function or array of numbers W(a, b) that describes the frequency content of f(t) for different values of t. A column W(ai, b) of this array (where i = 1, 2, . . . , n) lists the frequency spectrum of f(t) for a certain value (or range of values) of t. A row W(a, bi) contains numbers that describe how much of a certain frequency (or range of frequencies) f(t) has for any given t. The wavelet transform is such a method. It has been developed, researched, and applied to many areas of science and engineering since the early 1980s, although its roots 742 8. Wavelet Methods go much earlier. The main idea is to select a mother wavelet, a function that is nonzero in some small interval, and use it to explore the properties of f(t) in that interval. The mother wavelet is then translated to another interval of t and used again in the same way. Parameter b specifies the translation. Different frequency resolutions of f(t) are explored by scaling the mother wavelet with a scale factor a. Before getting into any details, we illustrate the relation between the “normal” (time domain) representation of a function and its two-dimensional transform by looking at the standard musical notation. This notation, used in the West for hundreds of years, is two-dimensional. Notes are written on a stave in two dimensions, where the horizontal axis denotes the time (from left to right) and the vertical axis denotes the pitch. The higher the note is on the stave, the higher the pitch of the tone played. In addition, the shape of a note indicates its duration. Figure 8.7a shows, from left to right, one whole note (called “C” in the U.S. and “do” in Europe), two half notes, three quarter notes, and two eighth notes. In addition to the stave and the notes, musical notation includes many other symbols and directions from the composer. However, the point is that the same music can also be represented by a one-dimensional function that simply describes the amplitude of the sound as a function of the time (Figure 8.7b). The two representations are mathematically equivalent, but are used differently in practice. The two-dimensional representation is used by musicians to actually perform the music. The one-dimensional representation is used to replay music that has already been performed and recorded. (a) (b) Time Amplitude Figure 8.7: Two Representations of Music.  Exercise 8.1: Come up with an example of a common notation that has a familiar two-dimensional representation used by humans and an unfamiliar one-dimensional representation used by machines. The principle of analyzing a function by time and by frequency can be applied to image compression because images contain areas that exhibit “trends” and areas with “anomalies.” A trend is an image feature that involves just a few frequencies (it is localized in frequency) but is spread spatially. A typical example is an image area where the brightness varies gradually. An anomaly is an image feature that involves several frequencies but is localized spatially (it is concentrated in a small image area). An example is an edge. We start by looking at functions that can serve as a wavelet, by defining the continuous wavelet transform (CWT) and its inverse, and illustrating the way the CWT works. We then show in detail how the Haar wavelet can be applied to the compression of images. This naturally leads to the concepts of filter banks and the discrete wavelet 8.5 The CWT and Its Inverse 743 transform (Section 8.8). The lifting scheme for the calculation of the wavelet transform and its inverse are described in Section 8.11. This is followed by descriptions of several compression methods that either employ the wavelet transform or compress the coefficients that result from such a transform. 8.5 The CWT and Its Inverse The continuous wavelet transform (CWT, [Lewalle 95] and [Rao and Bopardikar 98]) of a function f(t) involves a mother wavelet .(t). The mother wavelet can be any real or complex continuous function that satisfies the following properties: 1. The total area under the curve of the function is zero, i.e.,  . -. .(t) dt = 0. 2. The total area of |.(t)|2 is finite, i.e.  . -. |.(t)|2 dt < .. This condition implies that the integral of the square of the wavelet has to exist. We can also say that a wavelet has to be square integrable, or that it belongs to the set L2(R) of all the square integrable functions. 3. The admissibility condition, discussed below. Property 1 suggests a function that oscillates above and below the t axis. Such a function tends to have a wavy appearance. Property 2 implies that the energy of the function is finite, suggesting that the function is localized in some finite interval and is zero, or almost zero, outside this interval. These properties justify the name “wavelet.” An infinite number of functions satisfy these conditions, and some of them have been researched and are commonly used for wavelet transforms. Figure 8.8a shows the Morlet wavelet, defined by .(t) = e-t2 cos.t 2 ln 2. e-t2 cos(2.885.t). This is a cosine curve whose oscillations are dampened by the exponential (or Gaussian) factor. More than 99% of its energy is concentrated in the interval -2.5 . t . 2.5. Figure 8.8b shows the so-called Mexican hat wavelet, defined as .(t) = (1 - 2t2)e-t2 . This is the second derivative of the (negative) Gaussian function -0.5e-t2 . 744 8. Wavelet Methods The energy of a function A function y = f(x) relates each value of the independent variable x with a value of y. Plotting these pairs of values results in a representation of the function as a curve in two dimensions. The energy of a function is defined in terms of the area enclosed by this curve and the x axis. It makes sense to say that a curve that stays close to the axis has little energy, while a curve that spends “time” away from the x axis has more energy. Negative values of y push the curve under the x axis, where its area is considered negative, so the energy of f(t) is defined as the area under the curve of the nonnegative function f(x)2. If the function is complex, its absolute value is used in area calculations, so the energy of f(x) is defined as  . -. |f(x)|2dx. -5 5 -2 2 1 1 -2 2 -1 1 (a) (b) Figure 8.8: The (a) Morlet and (b) Mexican Hat Wavelets. Once a wavelet .(t) has been chosen, the CWT of a square integrable function f(t) is defined as W(a, b) =  . -. f(t) 1 |a| .* t - b a dt. (8.3) The transform is a function of the two real parameters a and b. The * denotes the complex conjugate. If we define a function .a,b(t) as .a,b(t) = 1 |a| . t - b a , 8.5 The CWT and Its Inverse 745 we can write Equation (8.3) in the form W(a, b) =  . -. f(t).a,b(t) dt. (8.4) Mathematically, the transform is the inner product of the two functions f(t) and .a,b(t). The quantity 1/|a| is a normalizing factor that ensures that the energy of .(t) remains independent of a and b, i.e.,  . -. |.a,b(t)|2 dt =  . -. |.(t)|2 dt. For any a, .a,b(t) is a copy of .a,0 shifted b units along the time axis. Thus, b is a translation parameter. Setting b = 0 shows that .a,0(t) = 1 |a| . t a , implying that a is a scaling (or a dilation) parameter. Values a > 1 stretch the wavelet, while values 0 < a < 1 shrink it. Since wavelets are used to transform a function, the inverse transform is needed. We denote by .(.) the Fourier transform of .(t): .(.) =  . -. .(t)e-i.t dt. If W(a, b) is the CWT of a function f(t) with a wavelet .(t), then the inverse CWT is defined by f(t) = 1 C  . -.  . -. 1 |a|2W(a, b).a,b(t) da db, where the quantity C is defined as C =  . -. |.(.)|2 |.| d.. The inverse CWT exists if C is positive and finite. Since C is defined by means of ., which itself is defined by means of the wavelet .(t), the requirement that C be positive and finite imposes another restriction, called the admissibility condition, on the choice of wavelet. The CWT is best thought of as an array of numbers that are inner products of f(t) and .a,b(t). The rows of the array correspond to values of a, and the columns are indexed by b. The inner product of two functions f(t) and g(t) is defined as f(t), g(t) =  . -. f(t)g*(t) dt, 746 8. Wavelet Methods so the CWT is the inner product f(t), .a,b(t) =  . -. f(t).a,b(t) dt. After this introduction, we are now in a position to explain the intuitive meaning of the CWT. We start with a simple example: the CWT of a sine wave, where the Mexican hat is chosen as the wavelet. Figure 8.9a shows a sine wave with two copies of the wavelet. Copy 1 is positioned at a point where the sine wave has a maximum. At this point there is a good match between the function being analyzed (the sine) and the wavelet. The wavelet replicates the features of the sine wave. As a result, the inner product of the sine and the wavelet is a large positive number. In contrast, copy 2 is positioned where the sine wave has a minimum. At that point the wave and the wavelet are almost mirror images of each other. Where the sine wave is positive, the wavelet is negative and vice versa. The product of the wave and the wavelet at this point is negative, and the CWT (the integral or the inner product) becomes a large negative number. In between points 1 and 2, the CWT drops from positive, to zero, to negative. As the wavelet is translated along the sine wave, from left to right, the CWT alternates between positive and negative and produces the small wave shown in Figure 8.9b. This shape is the result of the close match between the function being analyzed (the sine wave) and the analyzing wavelet. They have similar shapes and also similar frequencies. We now extend the analysis to cover different frequencies. This is done by scaling the wavelet. Figure 8.9c shows (in part 3) what happens when the wavelet is stretched such that it covers several periods of the sine wave. Translating the wavelet from left to right does not affect the match between it and the sine wave by much, with the result that the CWT varies just a little. The wider the wavelet, the closer the CWT is to a constant. Notice how the amplitude of the wavelet has been reduced, thereby reducing its area and producing a small constant. A similar thing happens in part 4 of the figure, where the wavelet has shrunk and is much narrower than one cycle of the sine wave. Since the wavelet is so “thin,” the inner product of it and the sine wave is always a small number (positive or negative) regardless of the precise position of the wavelet relative to the sine wave. We see that translating the wavelet does not much affect its match to the sine wave, resulting in a CWT that is close to a constant. The results of translating the wavelet, scaling it, and translating again and again are summarized in Figure 8.9d. This is a density plot of the transform function W(a, b) where the horizontal axis corresponds to values of b (translation) and the vertical axis corresponds to values of a (scaling). Figure 8.9e is a contour plot of the same W(a, b). These diagrams show that there is a good match between the function and the wavelet at a certain frequency (the frequency of the sine wave). At other frequencies the match deteriorates, resulting in a transform that gets closer and closer to a constant. This is how the CWT provides a time-frequency analysis of a function f(t). The result is a function W(a, b) of two variables that shows the match between f(t) and the wavelet at different frequencies of the wavelet and at different times. It is obvious that the quality of the match depends on the choice of wavelet. If the wavelet is very different from f(t) at any frequencies and any times, the values of the resulting W(a, b) will all be small and will not exhibit much variation. As a result, when wavelets are used to 8.5 The CWT and Its Inverse 747 50 100 150 200 250 5 10 15 20 25 30 50 100 150 200 250 5 10 15 20 25 30 (a) (b) Time (c) (d) (e) (1) (3) (4) (2) Figure 8.9: The Continuous Wavelet Transform of a Pure Sine Wave. t=linspace(0,6*pi,256); t=linspace(-10,10,256); sinwav=sin(t); sombr=(1-2*t.^2).*exp(-t.^2); plot(t,sinwav) plot(t,sombr) cwt=CWT(sinwav,10,’Sombrero’); axis(’ij’); colormap(gray); imagesc(cwt’) x=1:256; y=1:30; [X,Y]=meshgrid(x,y); contour(X,Y,cwt’,10) Code For Figure 8.9. 748 8. Wavelet Methods compress images, different wavelets should be selected for different image types (bi-level, continuous-tone, and discrete-tone), but the precise choice of wavelet is still the subject of much research. Since both our function and our wavelet are simple functions, it may be possible in this case to calculate the integral of the CWT as an indefinite integral, and come up with a closed formula for W(a, b). In the general case, however, this is impossible, either in practice or in principle, and the calculations have to be done numerically. The next example is slightly more complex and leads to a better understanding of the CWT. The function being analyzed this time is a sine wave with an accelerated frequency: the so-called chirp. It is given by f(t) = sin(t2), and it is shown in Figure 8.10a. The wavelet is the same Mexican hat. Readers who have gone over the previous example carefully will have no trouble in interpreting the CWT of this example. It is given by Figure 8.10b,c and shows how the frequency of f(t) increases with time. 50 100 150 200 250 5 10 15 20 25 50 100 150 200 250 30 5 10 15 20 25 30 (a) (b) (c) Figure 8.10: The Continuous Wavelet Transform of a Chirp. These two examples illustrate how the CWT is used to analyze the time frequency of a function f(t). It is clear that the user needs experience both in order to select the right wavelet for the job and in order to interpret the results. However, with the right experience, the CWT can be a powerful analytic tool in the hands of scientists and engineers.  Exercise 8.2: Experiment with the CWT by trying to work out the following analysis. Select the function f(t) = 1 + sin(.t + t2)/8 as your candidate for analysis. This is a 8.6 The Haar Transform 749 sine wave that oscillates between y = 1- 1/8 and y = 1+1/8 with a frequency 2. + t that increases with t. As the wavelet, select the Mexican hat. Plot several translated copies of the wavelet against f(t), then use appropriate software, such as Mathematica or Matlab, to calculate and display the CWT. Wavelets are also subject to the uncertainty principle (Section 8.3). The Haar wavelet is very well localized in the time domain but is spread in the frequency domain due to the appearance of sidebands in the Fourier spectrum. In contrast, the Mexican hat wavelet and especially the Morlet wavelet have more concentrated frequencies but are spread over time. An important result of the uncertainty principle is that it is impossible to achieve a complete simultaneous mapping of both time and frequency. Wavelets provide a compromise, or a near optimal solution, to this problem, and this is one feature that makes them superior to Fourier analysis. We can compare wavelet analysis to looking at a complex object first from a distance, then closer, then through a magnifying glass, and finally through a microscope. When looking from a distance, we see the overall shape of the object, but not any small details. When looking through a microscope, we see small details, but not the overall shape. This is why it is important to analyze in different scales. When we change the scale of the wavelet, we get new information about the function being analyzed. 8.6 The Haar Transform Alfr΄ed Haar (1885–1933) Alfr΄ed Haar was born in Budapest and received his higher mathematical training in G¨ottingen, where he later became a privatdozent. In 1912, he returned to Hungary and became a professor of mathematics first in Kolozsv΄ar and then in Szeged, where he and his colleagues created a major mathematical center. Haar is best remembered for his work on analysis on groups. In 1932 he introduced an invariant measure on locally compact groups, now called the Haar measure, which allows an analog of Lebesgue integrals to be defined on locally compact topological groups. Mathematical lore has it that John von Neumann tried to discourage Haar in this work because he felt certain that no such measure could exist. The following limerick celebrates Haar’s achievement. Said a mathematician named Haar, “Von Neumann can’t see very far. He missed a great treasure— They call it Haar measure— Poor Johnny’s just not up to par.” 750 8. Wavelet Methods Information that is being wavelet transformed in practical applications, such as digitized sound and images, is discrete, consisting of individual numbers. This is why the discrete, and not the continuous, wavelet transform is used in practice. The discrete wavelet transform is described in general in Section 8.8, but we precede this discussion by presenting a simple example of this type of transform, namely, the Haar wavelet transform. The use of the Haar transform for image compression is described here from a practical point of view. We first show how this transform is applied to the compression of grayscale images, then show how this method can be extended to color images. The Haar transform [Stollnitz et al. 96] was introduced in Section 7.7.3. The Haar transform uses a scale function .(t) and a wavelet .(t), both shown in Figure 8.11a, to represent a large number of functions f(t). The representation is the infinite sum f(t) = . k=-. ck.(t - k) + . k=-. . j=0 dj,k.(2jt - k), where ck and dj,k are coefficients to be calculated. The basic scale function .(t) is the unit pulse .(t) = 1, 0 . t < 1, 0, otherwise. The function .(t - k) is a copy of .(t), shifted k units to the right. Similarly, .(2t - k) is a copy of .(t - k) scaled to half the width of .(t - k). The shifted copies are used to approximate f(t) at different times t. The scaled copies are used to approximate f(t) at higher resolutions. Figure 8.11b shows the functions .(2jt - k) for j = 0, 1, 2, and 3 and for k = 0, 1, . . . , 7. The basic Haar wavelet is the step function .(t) =  1, 0 . t < 0.5, -1, 0.5 . t < 1. From this we can see that the general Haar wavelet .(2jt - k) is a copy of .(t) shifted k units to the right and scaled such that its total width is 1/2j .  Exercise 8.3: Draw the four Haar wavelets .(22t - k) for k = 0, 1, 2, and 3. Both .(2jt-k) and .(2jt-k) are nonzero in an interval of width 1/2j . This interval is their support. Since this interval tends to be short, we say that these functions have compact support. We illustrate the basic transform on the simple step function f(t) = 5, 0 . t < 0.5, 3, 0.5 . t < 1. It is easy to see that f(t) = 4.(t)+.(t). We say that the original steps (5, 3) have been transformed to the (low resolution) average 4 and the (high resolution) detail 1. Using 8.6 The Haar Transform 751 (a) 1 1 1 0 0 1 1 2 2 3 3 4 5 6 7 1 (b) t t .(t) .(2jt-k) (c) .(4t-k) .(t) j= k Figure 8.11: The Haar Basis Scale and Wavelet Functions. matrix notation, this can be expressed (up to a factor of .2) as (5, 3)A2 = (4, 1), where A2 is the order-2 Haar transform matrix of Equation (7.12). An image is a two-dimensional array of pixel values. To illustrate how the Haar transform is used to compress an image, we start with a single row of pixel values, i.e., a one-dimensional array of n values. For simplicity we assume that n is a power of 2. (We use this assumption throughout this chapter, but there is no loss of generality. If n has a different value, the data can be extended by appending zeros. After decompression, the extra zeros are removed.) Consider the array of eight values (1, 2, 3, 4, 5, 6, 7, 8). We first compute the four averages (1 + 2)/2 = 3/2, (3 + 4)/2 = 7/2, (5 + 6)/2 = 11/2, and (7 + 8)/2 = 15/2. It is impossible to reconstruct the original eight values from these four averages, so we also compute the four differences (1 - 2)/2 = -1/2, (3 - 4)/2 = -1/2, (5 - 6)/2 = -1/2, and (7 - 8)/2 = -1/2. These differences are called detail coefficients, and in this section the terms “difference” and “detail” are used interchangeably. We can think of the averages as a coarse resolution representation of the original image, and of the details as the data needed to reconstruct the original image from this coarse resolution. If the pixels of the image are correlated, the coarse representation will resemble the original pixels, while the details will be small. This explains why the Haar wavelet compression of images uses averages and details. It is easy to see that the array (3/2, 7/2, 11/2, 15/2,-1/2,-1/2,-1/2,-1/2) made of the four averages and four differences can be used to reconstruct the original eight 752 8. Wavelet Methods Prolonged, lugubrious stretches of Sunday afternoon in a university town could be mitigated by attending Sillery’s tea parties, to which anyone might drop in after half-past three. Action of some law of averages always regulated numbers at these gatherings to something between four and eight persons, mostly undergraduates, though an occasional don was not unknown. —Anthony Powell, A Question of Upbringing values. This array has eight values, but its last four components, the differences, tend to be small numbers, which helps in compression. Encouraged by this, we repeat the process on the four averages, the large components of our array. They are transformed into two averages and two differences, yielding (10/4, 26/4,-4/4,-4/4,-1/2,-1/2,-1/2,-1/2). The next, and last, iteration of this process transforms the first two components of the new array into one average (the average of all eight components of the original array) and one difference (36/8,-16/8,-4/4,-4/4,-1/2,-1/2,-1/2,-1/2). The last array is the Haar wavelet transform of the original data items. Because of the differences, the wavelet transform tends to have numbers smaller than the original pixel values, so it is easier to compress using RLE, perhaps combined with move-to-front and Huffman coding. Lossy compression can be obtained if some of the smaller differences are quantized or even completely deleted (changed to zero). Before we continue, it is interesting (and also useful) to calculate the complexity of this transform, i.e., the number of necessary arithmetic operations as a function of the size of the data. In our example we needed 8 + 4 + 2 = 14 operations (additions and subtractions), a number that can also be expressed as 14 = 2(8 - 1). In the general case, assume that we start with N = 2n data items. In the first iteration we need 2n operations, in the second one we need 2n-1 operations, and so on, until the last iteration, where 2n-(n-1) = 21 operations are needed. Thus, the total number of operations is n i=1 2i = n i=0 2i -1 = 1 - 2n+1 1 - 2 - 1 = 2n+1 - 2 = 2(2n - 1) = 2(N - 1). The Haar wavelet transform of N data items can therefore be performed with 2(N -1) operations, so its complexity is O(N), an excellent result. It is useful to associate with each iteration a quantity called resolution, which is defined as the number of remaining averages at the end of the iteration. The resolutions after each of the three iterations above are 4(= 22), 2(= 21), and 1(= 20). Section 8.6.3 shows that each component of the wavelet transform should be normalized by dividing it by the square root of the resolution. (This is the orthonormal Haar transform, also discussed in Section 7.7.3.) Thus, our example wavelet transform becomes 36/8 .20 , -16/8 .20 , -4/4 .21 , -4/4 .21 , -1/2 .22 , -1/2 .22 , -1/2 .22 , -1/2 .22 . If the normalized wavelet transform is used, it can be formally proved that ignoring the 8.6 The Haar Transform 753 smallest differences is the best choice for lossy wavelet compression, since it causes the smallest loss of image information. The two procedures of Figure 8.12 illustrate how the normalized wavelet transform of an array of n components (where n is a power of 2) can be computed. Reconstructing the original array from the normalized wavelet transform is illustrated by the pair of procedures of Figure 8.13. These procedures seem at first to be different from the averages and differences discussed earlier. They don’t compute averages, because they divide by .2 instead of by 2; the first procedure starts by dividing the entire array by .n, and the second one ends by doing the reverse. The final result, however, is the same as that shown above. Starting with array (1, 2, 3, 4, 5, 6, 7, 8), the three iterations of procedure NWTcalc result in the following 3 .24 , 7 .24 , 11 .24 , 15 .24 , -1 .24 , -1 .24 , -1 .24 , -1 .24 , 10 .25 , 26 .25 , -4 .25 , -4 .25 , -1 .24 , -1 .24 , -1 .24 , -1 .24 , 36 .26 , -16 .26 , -4 .25 , -4 .25 , -1 .24 , -1 .24 , -1 .24 , -1 .24 , 36/8 .20 , -16/8 .20 , -4/4 .21 , -4/4 .21 , -1/2 .22 , -1/2 .22 , -1/2 .22 , -1/2 .22 . We spake no word, Tho’ each I ween did hear the other’s soul. Not a wavelet stirred, And yet we heard The loneliest music of the weariest waves That ever roll. —Abram J. Ryan, Poems 8.6.1 Applying the Haar Transform Once the concept of a wavelet transform is grasped, it’s easy to generalize it to a complete two-dimensional image. This can be done in several ways that are discussed in Section 8.10. Here we show two such approaches, called the standard decomposition and the pyramid decomposition. The former (Figure 8.15) starts by computing the wavelet transform of every row of the image. This results in a transformed image where the first column contains averages and all the other columns contain differences. The standard algorithm then computes the wavelet transform of every column. This results in one average value at the top-left corner, with the rest of the top row containing averages of differences, and with all other pixel values transformed into differences. 754 8. Wavelet Methods procedure NWTcalc(a:array of real, n:int); comment n is the array size (a power of 2) a:=a/.n comment divide entire array j:=n; while j. 2 do NWTstep(a, j); j:=j/2; endwhile; end; procedure NWTstep(a:array of real, j:int); for i=1 to j/2 do b[i]:=(a[2i-1]+a[2i])/.2; b[j/2+i]:=(a[2i-1]-a[2i])/.2; endfor; a:=b; comment move entire array end; Figure 8.12: Computing the Normalized Wavelet Transform. procedure NWTreconst(a:array of real, n:int); j:=2; while j.n do NWTRstep(a, j); j:=2j; endwhile a:=a.n; comment multiply entire array end; procedure NWTRstep(a:array of real, j:int); for i=1 to j/2 do b[2i-1]:=(a[i]+a[j/2+i])/.2; b[2i]:=(a[i]-a[j/2+i])/.2; endfor; a:=b; comment move entire array end; Figure 8.13: Restoring From a Normalized Wavelet Transform. 8.6 The Haar Transform 755 The latter method computes the wavelet transform of the image by alternating between rows and columns. The first step is to calculate averages and differences for all the rows (just one iteration, not the entire wavelet transform). This creates averages in the left half of the image and differences in the right half. The second step is to calculate averages and differences (just one iteration) for all the columns, which results in averages in the top-left quadrant of the image and differences elsewhere. Steps 3 and 4 operate on the rows and columns of that quadrant, resulting in averages concentrated in the top-left subquadrant. Pairs of steps are repeatedly executed on smaller and smaller subsquares, until only one average is left, at the top-left corner of the image, and all other pixel values have been reduced to differences. This process is summarized in Figure 8.16. The transforms described in Section 7.6 are orthogonal. They transform the original pixels into a few large numbers and many small numbers. In contrast, wavelet transforms, such as the Haar transform, are subband transforms. They partition the image into regions such that one region contains large numbers (averages in the case of the Haar transform) and the other regions contain small numbers (differences). However, these regions, which are called subbands, are more than just sets of large and small numbers. They reflect different geometrical artifacts of the image. To illustrate this important feature, we examine a small, mostly-uniform image with one vertical line and one horizontal line. Figure 8.14a shows an 8 Χ 8 image with pixel values of 12, except for a vertical line with pixel values of 14 and a horizontal line with pixel values of 16. 12 12 12 12 14 12 12 12 12 12 12 12 14 12 12 12 12 12 12 12 14 12 12 12 12 12 12 12 14 12 12 12 12 12 12 12 14 12 12 12 16 16 16 16 14 16 16 16 12 12 12 12 14 12 12 12 12 12 12 12 14 12 12 12 12 12 13 12 0 0 2 0 12 12 13 12 0 0 2 0 12 12 13 12 0 0 2 0 12 12 13 12 0 0 2 0 12 12 13 12 0 0 2 0 16 16 15 16 0 0 2 0 12 12 13 12 0 0 2 0 12 12 13 12 0 0 2 0 12 12 13 12 0 0 2 0 12 12 13 12 0 0 2 0 14 14 14 14 0 0 0 0 12 12 13 12 0 0 2 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 4 4 2 4 0 0 4 0 0 0 0 0 0 0 0 0 (a) (b) (c) Figure 8.14: An 8Χ8 Image and Its Subband Decomposition. Figure 8.14b shows the results of applying the Haar transform once to the rows of the image. The right half of this figure (the differences) is mostly zeros, reflecting the uniform nature of the image. However, traces of the vertical line can easily be seen (the notation 2 indicates a negative difference). Figure 8.14c shows the results of applying the Haar transform once to the columns of Figure 8.14b. The upper-right subband now features traces of the vertical line, whereas the lower-left subband shows traces of the horizontal line. These subbands are denoted by HL and LH, respectively (Figures 8.16 and 8.55, although there is inconsistency in the use of this notation by various authors). The lower-right subband, denoted by HH, reflects diagonal image artifacts (which our example image lacks). Most interesting is the upper-left subband, denoted by LL, that consists entirely of averages. This subband is a one-quarter version of the entire image, containing traces of both the vertical and the horizontal lines. 756 8. Wavelet Methods procedure StdCalc(a:array of real, n:int); comment array size is nxn (n = power of 2) for r=1 to n do NWTcalc(row r of a, n); endfor; for c=n to 1 do comment loop backwards NWTcalc(col c of a, n); endfor; end; procedure StdReconst(a:array of real, n:int); for c=n to 1 do comment loop backwards NWTreconst(col c of a, n); endfor; for r=1 to n do NWTreconst(row r of a, n); endfor; end; Original image L1 H1 L2 H2 H1 H2 H1 L3 H3 Figure 8.15: The Standard Image Wavelet Transform and Decomposition. 8.6 The Haar Transform 757 procedure NStdCalc(a:array of real, n:int); a:=a/.n comment divide entire array j:=n; while j. 2 do for r=1 to j do NWTstep(row r of a, j); endfor; for c=j to 1 do comment loop backwards NWTstep(col c of a, j); endfor; j:=j/2; endwhile; end; procedure NStdReconst(a:array of real, n:int); j:=2; while j.n do for c=j to 1 do comment loop backwards NWTRstep(col c of a, j); endfor; for r=1 to j do NWTRstep(row r of a, j); endfor; j:=2j; endwhile a:=a.n; comment multiply entire array end; Original image L H HL HH LL LH HL HH LH LLL LLH HL HH LH Figure 8.16: The Pyramid Image Wavelet Transform. 758 8. Wavelet Methods  Exercise 8.4: Construct a diagram similar to Figure 8.14 to show how subband HH reflects diagonal artifacts of the image. (Artifact: A feature not naturally present, introduced during preparation or investigation.) Figure 8.55 shows four levels of subbands, where level 1 contains the detailed features of the image (also referred to as the high-frequency or fine-resolution wavelet coefficients) and the top level, level 4, contains the coarse image features (low-frequency or coarse-resolution coefficients). It is clear that the lower levels can be quantized coarsely without much loss of important image information, while the higher levels should be quantized finely. The subband structure is the basis of all the image compression methods that use the wavelet transform. Figure 8.17 shows typical results of the pyramid wavelet transform. The original image is shown in Figure 8.17a, and Figure 8.17c is a general pyramid decomposition. In order to illustrate how the pyramid transform works, this image consists only of horizontal, vertical, and slanted lines. The four quadrants of Figure 8.17b show smaller versions of the image. The top-left subband, containing the averages, is similar to the entire image, while each of the other three subbands shows image details. Because of the way the pyramid transform is constructed, the top-right subband contains vertical details, the bottom-left subband contains horizontal details, and the bottom-right one contains the details of slanted lines. Figure 8.17c shows the results of repeatedly applying this transform. The image is transformed into subbands of horizontal, vertical, and diagonal details, while the top-left subsquare, containing the averages, is shrunk to a single pixel. Section 7.6 discusses image transforms. It should be mentioned that there are two main types of transforms, orthogonal and subband. An orthogonal linear transform is performed by computing the inner product of the data (pixel values or sound samples) with a set of basis functions. The result is a set of transform coefficients that can later be quantized or compressed with RLE, Huffman coding, or other methods. Several examples of important orthogonal transforms, such as the DCT, the WHT, and the KLT, are described in detail in Section 7.6. The Fourier transform also belongs in this category. It is discussed in Section 8.1. The other main type of transform is the subband transform. It is performed by computing a convolution of the data (Section 8.7) with a set of bandpass filters. Each resulting subband encodes a particular portion of the frequency content of the data. As a reminder, the discrete inner product of the two vectors fi and gi is defined by f, g = i fi gi. The discrete convolution h is defined by Equation (8.5): hi = f  g = j fj gi-j . (8.5) (Each element hi of the discrete convolution h is the sum of products. It depends on i in the special way shown.) 8.6 The Haar Transform 759 (a) (c) (b) Figure 8.17: An Example of the Pyramid Image Wavelet Transform. 760 8. Wavelet Methods Either method, standard or uniform, results in a transformed, although not yet compressed, image that has one average at the top-left corner and smaller numbers, differences, or averages of differences everywhere else. These numbers can be compressed using a combination of methods, such as RLE, move-to-front, and Huffman coding. If lossy compression is acceptable, some of the smallest differences can be quantized or even set to zeros, which creates run lengths of zeros, making the use of RLE even more attractive. Whiter foam than thine, O wave, Wavelet never wore, Stainless wave; and now you lave The far and stormless shore — Ever — ever — evermore! —Abram J. Ryan, Poems Color Images: So far we have assumed that each pixel is a single number (i.e., we have a single-component image, in which all pixels are shades of the same color, normally gray). Any compression method for single-component images can be extended to color (three-component) images by separating the three components, then transforming and compressing each individually. If the compression method is lossy, it makes sense to convert the three image components from their original color representation, which is normally RGB, to the YIQ color representation. The Y component of this representation is called luminance, and the I and Q (the chrominance) components are responsible for the color information [Salomon 99]. The advantage of this color representation is that the human eye is most sensitive to Y and least sensitive to Q. A lossy method should therefore leave the Y component alone and delete some data from the I, and more data from the Q components, resulting in good compression and in a loss to which the eye is not that sensitive. It is interesting to note that U.S. color television transmission also takes advantage of the YIQ representation. Signals are broadcast with bandwidths of 4 MHz for Y, 1.5 MHz for I, and only 0.6 MHz for Q. 8.6.2 Properties of the Haar Transform The examples in this section illustrate some properties of the Haar transform, and of the discrete wavelet transform in general. Figure 8.18 shows a highly correlated 8Χ8 image and its Haar wavelet transform. Both the grayscale and numeric values of the pixels and the transform coefficients are shown. Because the original image is so correlated, the wavelet coefficients are small and there are many zeros.  Exercise 8.5: A glance at Figure 8.18 suggests that the last sentence is wrong. The wavelet transform coefficients listed in the figure are very large compared with the pixel values of the original image. In fact, we know that the top-left Haar transform coefficient should be the average of all the image pixels. Since the pixels of our image have values that are (more or less) uniformly distributed in the interval [0, 255], this average should be around 128, yet the top-left transform coefficient is 1051. Explain this! 8.6 The Haar Transform 761 255 224 192 159 127 95 63 32 0 32 64 159 127 159 191 223 255 224 192 159 127 95 63 32 0 32 64 159 127 159 191 223 255 224 192 159 127 95 63 32 0 32 64 159 127 159 191 223 255 224 192 159 127 95 63 32 0 32 64 159 127 159 191 223 1051 34.0 -44.5 -0.7 -1.0 -62 0 -1.0 0 0.0 0.0 0.0 0.0 0 0 0.0 0 0.0 0.0 0.0 0.0 0 0 0.0 0 0.0 0.0 0.0 0.0 0 0 0.0 48 239.5 112.8 90.2 31.5 64 32 31.5 48 239.5 112.8 90.2 31.5 64 32 31.5 48 239.5 112.8 90.2 31.5 64 32 31.5 48 239.5 112.8 90.2 31.5 64 32 31.5 Figure 8.18: The 8Χ8 Image Reconstructed in Figure 8.21 and Its Haar Transform. (a) (b) Figure 8.19: (a) A 128Χ128 Image With Activity on the Right. (b) Its Transform. 762 8. Wavelet Methods In a discrete wavelet transform, most of the wavelet coefficients are details (or differences). The details in the lower levels represent the fine details of the image. As we move higher in the subband level, we find details that correspond to coarser image features. Figure 8.19a illustrates this concept. It shows an image that is smooth on the left and has “activity” (i.e., adjacent pixels that tend to be different) on the right. Part (b) shows the wavelet transform of the image. Low levels (corresponding to fine details) have transform coefficients on the right, since this is where the image activity is located. High levels (coarse details) look similar but also have coefficients on the left side, because the image is not completely blank on the left. The Haar transform is the simplest wavelet transform, but even this simple method illustrates the power of the wavelet transform. It turns out that the low levels of the discrete wavelet transform contain the unimportant image features, so quantizing or discarding these coefficients can lead to lossy compression that is both efficient and of high quality. Often, the image can be reconstructed from very few transform coefficients without any noticeable loss of quality. Figure 8.21a–c shows three reconstructions of the simple 8Χ8 image of Figure 8.18. They were obtained from only 32, 13, and 5 wavelet coefficients, respectively. 0 20 40 60 80 100 120 0 20 40 60 80 100 120 nz = 653 (a) (b) nz=653 0 20 40 60 80 100 120 0 20 40 60 80 100 120 Figure 8.20: Reconstructing a 128Χ128 Simple Image From 4% of Its Coefficients. Figure 8.20 is a similar example. It shows a bi-level image fully reconstructed from just 4% of its transform coefficients (653 coefficients out of 128Χ128). Experimenting is the key to understanding these concepts. Proper mathematical software makes it easy to input images and experiment with various features of the discrete wavelet transform. In order to help the interested reader, Figure 8.22 lists a Matlab program that inputs an image, computes its Haar wavelet transform, discards a given percentage of the smallest transform coefficients, then computes the inverse transform to reconstruct the image. 8.6 The Haar Transform 763 0 1 2 3 4 5 6 7 8 9 1 2 3 4 5 6 7 8 nz = 32 0 1 2 3 4 5 6 7 8 9 1 2 3 4 5 6 7 8 nz = 13 0 1 2 3 4 5 6 7 8 9 1 2 3 4 5 6 7 8 nz = 5 (a) (b) (c) nz=5 0 1 2 3 4 5 6 7 8 9 nz=13 0 1 2 3 4 5 6 7 8 9 nz=32 0 1 2 3 4 5 6 7 8 9 1 2 3 4 5 6 7 8 1 2 3 4 5 6 7 8 1 2 3 4 5 6 7 8 Figure 8.21: Three Lossy Reconstructions of an 8Χ8 Image. 764 8. Wavelet Methods Lossy wavelet image compression involves the discarding of coefficients, so the concept of sparseness ratio is defined to measure the amount of coefficients discarded. Sparseness is defined as the number of nonzero wavelet coefficients divided by the number of coefficients left after some are discarded. The higher the sparseness ratio, the fewer coefficients are left. Higher sparseness ratios lead to better compression but may result in poorly reconstructed images. The sparseness ratio is distantly related to compression factor, a compression measure defined in the Introduction. The line “filename=’lena128’; dim=128;” contains the image file name and the dimension of the image. The image files used by the authors were in raw form and contained just the grayscale values, each as a single byte. There is no header, and not even the image resolution (number of rows and columns) is included in the file. However, Matlab can read other types of files. The image is assumed to be square, and parameter “dim” should be a power of 2. The assignment “thresh=” specifies the percentage of transform coefficients to be deleted. This provides an easy way to experiment with lossy wavelet image compression. File “harmatt.m” contains two functions that compute the Haar wavelet coefficients in a matrix form (Section 8.6.3). (A technical note: A Matlab m file may include commands or a function but not both. It may, however, contain more than one function, provided that only the top function is invoked from outside the file. All the other functions must be called from within the file. In our case, function harmatt(dim) calls function individ(n).)  Exercise 8.6: Use the code of Figure 8.22 (or similar code) to compute the Haar transform of the “Lena” image (Figure 7.55) and reconstruct it three times by discarding more and more detail coefficients. 8.6.3 A Matrix Approach The principle of the Haar transform is to compute averages and differences. It turns out that this can be done by means of matrix multiplication ([Mulcahy 96] and [Mulcahy 97]). As an example, we look at the top row of the simple 8Χ8 image of Figure 8.18. Anyone with a little experience with matrices can construct a matrix that when multiplied by this vector creates a vector with four averages and four differences. Matrix A1 of Equation (8.6) does that and, when multiplied by the top row of pixels of Figure 8.18, generates (239.5, 175.5, 111.0, 47.5, 15.5, 16.5, 16.0, 15.5). Similarly, matrices A2 and A3 perform the second and third steps of the transform, respectively. The results are shown in Equation (8.7): A1 = ........... 1 2 1 2 0 0 0 0 0 0 0 0 1 2 1 2 0 0 0 0 0 0 0 0 1 2 1 2 0 0 0 0 0 0 0 0 1 2 1 2 1 2 -1 2 0 0 0 0 0 0 0 0 1 2 -1 2 0 0 0 0 0 0 0 0 1 2 -1 2 0 0 0 0 0 0 0 0 1 2 -1 2 ........... , A1 ........... 255 224 192 159 127 95 63 32 ........... = ........... 239.5 175.5 111.0 47.5 15.5 16.5 16.0 15.5 ........... , (8.6) 8.6 The Haar Transform 765 clear; % main program filename=’lena128’; dim=128; fid=fopen(filename,’r’); if fid==-1 disp(’file not found’) else img=fread(fid,[dim,dim])’; fclose(fid); end thresh=0.0; % percent of transform coefficients deleted figure(1), imagesc(img), colormap(gray), axis off, axis square w=harmatt(dim); % compute the Haar dim x dim transform matrix timg=w*img*w’; % forward Haar transform tsort=sort(abs(timg(:))); tthresh=tsort(floor(max(thresh*dim*dim,1))); cim=timg.*(abs(timg) > tthresh); [i,j,s]=find(cim); dimg=sparse(i,j,s,dim,dim); % figure(2) displays the remaining transform coefficients %figure(2), spy(dimg), colormap(gray), axis square figure(2), image(dimg), colormap(gray), axis square cimg=full(w’*sparse(dimg)*w); % inverse Haar transform density = nnz(dimg); disp([num2str(100*thresh) ’% of smallest coefficients deleted.’]) disp([num2str(density) ’ coefficients remain out of ’ ... num2str(dim) ’x’ num2str(dim) ’.’]) figure(3), imagesc(cimg), colormap(gray), axis off, axis square File harmatt.m with two functions function x = harmatt(dim) num=log2(dim); p = sparse(eye(dim)); q = p; i=1; while i<=dim/2; q(1:2*i,1:2*i) = sparse(individ(2*i)); p=p*q; i=2*i; end x=sparse(p); function f=individ(n) x=[1, 1]/sqrt(2); y=[1,-1]/sqrt(2); while min(size(x)) < n/2 x=[x, zeros(min(size(x)),max(size(x)));... zeros(min(size(x)),max(size(x))), x]; end while min(size(y)) < n/2 y=[y, zeros(min(size(y)),max(size(y)));... zeros(min(size(y)),max(size(y))), y]; end f=[x;y]; Figure 8.22: Matlab Code for the Haar Transform of an Image. 766 8. Wavelet Methods A2 = ........... 1 2 1 2 0 0 0 0 0 0 0 0 1 2 1 2 0 0 0 0 1 2 -1 2 0 0 0 0 0 0 0 0 1 2 -1 2 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 1 ........... , A3 = ............ 1 2 1 2 0 0 0 0 0 0 1 2 -1 2 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 1 ............ , A2 ........... 239.5 175.5 111.0 47.5 15.5 16.5 16.0 15.5 ........... = ........... 207.5 79.25 32.0 31.75 15.5 16.5 16.0 15.5 ........... , A3 ........... 207.5 79.25 32.0 31.75 15.5 16.5 16.0 15.5 ........... = ........... 143.375 64.125 32. 31.75 15.5 16.5 16. 15.5 ........... . (8.7) Instead of calculating averages and differences, all we have to do is construct matrices A1, A2, and A3, multiply them to get W = A1A2A3, and apply W to all the columns of the image I by multiplying W·I: W ........... 255 224 192 159 127 95 63 32 ........... = ............ 1 8 1 8 1 8 1 8 1 8 1 8 1 8 1 8 1 8 1 8 1 8 1 8 -1 8 -1 8 -1 8 -1 8 1 4 1 4 -1 4 -1 4 0 0 0 0 0 0 0 0 1 4 1 4 -1 4 -1 4 1 2 -1 2 0 0 0 0 0 0 0 0 1 2 -1 2 0 0 0 0 0 0 0 0 1 2 -1 2 0 0 0 0 0 0 0 0 1 2 -1 2 ............ ........... 255 224 192 159 127 95 63 32 ........... = ........... 143.375 64.125 32 31.75 15.5 16.5 16 15.5 ........... . This, of course, is only half the job. In order to compute the complete transform, we still have to apply W to the rows of the product W·I, and we do this by applying it to the columns of the transpose (W·I)T , then transposing the result. Thus, the complete transform is (see line timg=w*img*w’ in Figure 8.22) Itr = W(W·I)T T = W·I ·WT . The inverse transform is performed by W-1(W-1 ·IT tr)T = W-1Itr ·(W-1)T , and this is where the normalized Haar transform (mentioned on page 752) becomes important. Instead of calculating averages [quantities of the form (di + di+1)/2] and differences [quantities of the form (di - di+1)/2], it is better to use the quantities (di + 8.7 Filter Banks 767 di+1)/.2 and (di - di+1)/.2. This results is an orthonormal matrix W, and it is well known that the inverse of such a matrix is simply its transpose. Thus, we can write the inverse transform in the simple formWT·Itr·W [see line cimg=full(w’*sparse(dimg)*w) in Figure 8.22]. In between the forward and inverse transforms, some transform coefficients may be quantized or deleted. Alternatively, matrix Itr may be compressed by means of run length encoding and/or Huffman codes. Function individ(n) of Figure 8.22 starts with a 2Χ2 Haar transform matrix (notice that it uses .2 instead of 2) and then uses it to construct as many individual matrices Ai as necessary. Function harmatt(dim) combines those individual matrices to form the final Haar matrix for an image of dim rows and dim columns.  Exercise 8.7: Perform the calculation W·I ·WT for the 8Χ8 image of Figure 8.18. 8.7 Filter Banks This section uses the matrix approach to the Haar transform to introduce the reader to the idea of filter banks. We show how the Haar transform can be interpreted as a bank of two filters, a lowpass and a highpass. We explain the terms “filter,” “lowpass,” and “highpass” and show how the idea of filter banks leads naturally to the concept of subband transform. The Haar transform, of course, is the simplest wavelet transform, so it is used here to illustrate the new concepts. However, using it as a filter bank may not be very efficient. Most practical applications of wavelet filters employ more sophisticated sets of filter coefficients, but they are all based on the concept of filters and filter banks [Strang and Nguyen 96]. And just like the wavelet that moans on the beach, And, sighing, sinks back to the sea, So my song—it just touches the rude shores of speech, And its music melts back into me. —Abram J. Ryan, Poems A filter is a linear operator defined in terms of its filter coefficients h(0), h(1), h(2), . . .. It can be applied to an input vector x to produce an output vector y according to y(n) = k h(k)x(n - k) = h  x, where the symbol  indicates a convolution. Notice that the limits of the sum above have not been stated explicitly. They depend on the sizes of vectors x and h. Since our independent variable is the time t, it is convenient to assume that the inputs (and, consequently, also the outputs) come at all times t = . . . ,-2,-1, 0, 1, 2, . . .. Thus, we use the notation x = (. . . , a, b, c, d, e, . . .), 768 8. Wavelet Methods where the central value c is the input at time zero [c = x(0)], values d and e are the inputs at times 1 and 2, respectively, and b = x(-1) and a = x(-2). In practice, the inputs are always finite, so the infinite vector x will have only a finite number of nonzero elements. Deeper insight into the behavior of a linear filter can be gained by considering the simple input x = (. . . , 0, 0, 1, 0, 0, . . .). This input is zero at all times except at t = 0. It is called a unit pulse or a unit impulse. Even though the limits of the sum in the convolution have not been specified, it is easy to see that for any n there is only one nonzero term in the sum, so y(n) = h(n)x(0) = h(n). We say that the output y(n) = h(n) at time n is the response at time n to the unit impulse x(0) = 1. Since the number of filter coefficients h(i) is finite, this filter is a finite impulse response or FIR. Figure 8.23 shows the basic idea of a filter bank. It shows an analysis bank consisting of two filters, a lowpass filter H0 and a highpass filter H1. The lowpass filter employs convolution to remove the high frequencies from the input signal x and let the low frequencies through. The highpass filter does the opposite. Together, they separate the input into frequency bands. H0 H1 v2 v2 F0 F1 ^2 ^2 input x output x quantize compress or save Figure 8.23: A Two-Channel Filter Bank. The input x can be a one-dimensional signal (a vector of real numbers, which is what we assume in this section) or a two-dimensional signal, an image. The elements x(n) of x are fed into the filters one by one, and each filter computes and outputs one number y(n) in response to x(n). The number of responses is therefore double the number of inputs (because we have two filters); a bad result, since we are interested in data compression. To correct this situation, each filter is followed by a downsampling process where the odd-numbered outputs are thrown away. This operation is also called decimation and is represented by the boxes marked “v2”. After decimation, the number of outputs from the two filters together equals the number of inputs. Notice that the filter bank described here, followed by decimation, performs exactly the same calculations as matrix W = A1A2A3 of Section 8.6.3. Filter banks are just a more general way of looking at the Haar transform (or, in general, at the discrete wavelet transform). We look at this transform as a filtering operation, followed by decimation, and we can then try to find better filters. The reason for having a bank of filters as opposed to just one filter is that several filters working together, with downsampling, can exhibit behavior that is impossible to obtain with just a single filter. The most important feature of a filter bank is its ability to reconstruct the input from the outputs H0x and H1x, even though each has been decimated. Downsampling is not time invariant. After downsampling, the output is the evennumbered values y(0), y(2), y(4),. . . , but if we delay the inputs by one time unit, the new outputs will be y(-1), y(1), y(3),. . . , and these are different from and independent of the original outputs. These two sequences of signals are two phases of vector y. 8.7 Filter Banks 769 The outputs of the analysis bank are called subband coefficients. They can be quantized (if lossy compression is acceptable), and they can be compressed by means of RLE, Huffman, arithmetic coding, or any other method. Eventually, they are fed into the synthesis bank, where they are first upsampled (by inserting zeros for each oddnumbered coefficient that was thrown away), then passed through the inverse filters F0 and F1, and finally combined to form a single output vector ˆx. The output of each analysis filter (after decimation) is (vy) = (. . . , y(-4), y(-2), y(0), y(2), y(4), . . .). Upsampling inserts zeros for the decimated values, so it converts the output vector above to (^y) = (. . . , y(-4), 0, y(-2), 0, y(0), 0, y(2), 0, y(4), 0, . . .). Downsampling causes loss of data. Upsampling alone cannot compensate for it, because it simply inserts zeros for the missing data. In order to achieve lossless reconstruction of the original signal x, the filters have to be designed such that they compensate for this loss of data. One feature that is commonly used in the design of good filters is orthogonality. Figure 8.24 shows a set of orthogonal filters of size 4. The filters of the set are orthogonal because their dot product is zero (a, b, c, d) · (d,-c, b,-a) = 0. Notice how similar H0 and F0 are (and also H1 and F1). It still remains, of course, to choose actual values for the four filter coefficients a, b, c, and d. A full discussion of this is outside the scope of this book, but Section 8.7.1 illustrates some of the methods and rules used in practice to determine the values of various filter coefficients. An example is the Daubechies D4 filter, whose values are listed in Equation (8.11). v2 v2 ^2 ^2 x(n) x(n-3) d, -c, b,-a a, b, c, d d, c, b, a -a, b, -c, d Figure 8.24: An Orthogonal Filter Bank With Four Filter Coefficients. Simulating the operation of this filter manually shows that the reconstructed input is identical to the original input but lags three time units behind it. A filter bank can also be biorthogonal, a less restricted type of filter. Figure 8.25 shows an example of such a set of filters that can reconstruct a signal exactly. Notice the similarity of H0 and F1 and also of H1 and F0. v2 v2 ^2 ^2 x(n) 16 x(n-3) 1, -2, 1 -1, 2, 6, 2,-1 1, 2, 1 1, 2, -6, 2, 1 H0 H1 F0 F1 Figure 8.25: A Biorthogonal Filter Bank With Perfect Reconstruction. 770 8. Wavelet Methods We already know, from the discussion in Section 8.6, that the outputs of the lowpass filter H0 are normally passed through the analysis filter several times, creating shorter and shorter outputs. This recursive process can be illustrated as a tree (Figure 8.26). Since each node of this tree produces half the number of outputs as its predecessor, the tree is called a logarithmic tree. Figure 8.26 shows how the scaling function .(t) and the wavelet .(t) are obtained at the limit of the logarithmic tree. This is the connection between the discrete wavelet transform (using filter banks) and the CWT. v2 f(t) H1 Scaling function .(t) Wavelet .(t) H0 v2 H1 v2 H0 v2 H1 v2 H0 v2 Figure 8.26: Scaling Function and Wavelet as Limits of a Logarithmic Tree. As we “climb” up the logarithmic tree from level i to the next finer level i + 1, we compute the new averages from the new, higher-resolution scaling functions .(2it - k) and the new details from the new wavelets .(2it - k) signal at level i (averages)  + signal at level i + 1. details at level i (differences)  Each level of the tree corresponds to twice the frequency (or twice the resolution) of the preceding level, which is why the logarithmic tree is also called a multiresolution tree. Successive filtering through the tree separates lower and lower frequencies. People who do quantitative work with sound and music know that two tones at frequencies . and 2. sound like the same note and differ only by pitch. The frequency interval between . and 2. is divided into 12 subintervals (the so-called chromatic scale), but Western music has a tradition of favoring just eight of the twelve tones that result from this division (a diatonic scale, made up of seven notes, with the eighth note as the “octave”). This is why the basic frequency interval used in music is traditionally called an octave. We therefore say that adjacent levels of the multiresolution tree differ in an octave of frequencies. In order to understand the meaning of lowpass and highpass we need to work in the frequency domain, where the convolution of two vectors is replaced by a multiplication of their Fourier transforms. The vector x(n) is in the time domain, so its frequency domain equivalent is its discrete Fourier transform (Equation (8.1)) X(.) def = X(ei.) = . -. x(n)e-in., 8.7 Filter Banks 771 which is sometimes written in the z-domain, X(z) = . -. x(n)z-n, where z def = ei.. The convolution by h in the time domain now becomes a multiplication by the function H(.) = h(n)e-in. in the frequency domain, so we can express the output in the frequency domain by Y (ei.) = H(ei.)X(ei.), or, in reduced notation, Y (.) = H(.)X(.), or, in the z-domain, Y (z) = H(z)X(z). When all the inputs are X(.) = 1, the output at frequency . is Y (.) = H(.). We can now understand the operation of the lowpass Haar filter. It works by averaging two consecutive inputs, so it produces the output y(n) = 1 2x(n) + 1 2x(n - 1). (8.8) This is a convolution with only the two terms k = 0 and k = 1 in the sum. The filter coefficients are h(0) = h(1) = 1/2, and we can call the output a moving average, since each y(n) depends on the current input and its predecessor. If the input is the unit impulse x = (. . . , 0, 0, 1, 0, 0, . . .), then the output is y(0) = y(1) = 1/2, or y = (. . . , 0, 0, 1/2, 1/2, 0, . . .). The output values are simply the filter coefficients as we saw earlier. We can look at this averaging filter as the combination of an identity operator and a delay operator. The output produced by the identity operator equals the current input, while the output produced by the delay is the input one time unit earlier. Thus, we write averaging filter = 1 2 (identity) + 1 2 (delay). In matrix notation this can be expressed by ..... · · · y(-1) y(0) y(1) · · · .....=...... · · · 1 2 1 21 2 1 21 2 1 2 · · · ...... ..... · · · x(-1) x(0) x(1) · · · ..... . The 1/2 values on the main diagonal are copies of the weight of the identity operator. They all equal to the h(0) Haar filter coefficient. The 1/2 values on the diagonal below are copies of the weights of the delay operator. They all equal the h(1) Haar filter coefficient. Thus, the matrix is a constant diagonal matrix (or a banded matrix). A wavelet filter that has a coefficient h(3) would correspond to a matrix where this filter 772 8. Wavelet Methods coefficient appears on the second diagonal below the main diagonal. The rule of matrix multiplication produces the familiar convolution y(n) = h(0)x(n) + h(1)x(n - 1) + h(2)x(n - 2) + · · · = k h(k)x(n - k). Notice that the matrix is lower triangular. The upper diagonal, which would naturally correspond to the filter coefficients h(-1), h(-2),. . . , is zero. All filter coefficients with negative indices must be zero, since such coefficients lead to outputs that precede the inputs in time. In the real world, we are used to a cause preceding its effect, so our finite impulse response filters should also be causal. Summary: A causal FIR filter with N +1 filter coefficients h(0), h(1), . . . , h(N) (a filter with N+1 “taps”) has h(i) = 0 for all negative i and for i > N. When expressed in terms of a matrix, the matrix is lower triangular and banded. Such filters are commonly used and are important. From the Dictionary Tap (noun). 1. A cylindrical plug or stopper for closing an opening through which liquid is drawn, as in a cask; spigot. 2. A faucet or cock. 3. A connection made at an intermediate point on an electrical circuit or device. 4. An act or instance of wiretapping. To illustrate the frequency response of a filter we select an input vector of the form x(n) = ein. = cos(n.) + i sin(n.), for -.< n < .. This is a complex function whose real and imaginary parts are a cosine and a sine, respectively, both with frequency .. Recall that the Fourier transform of a pulse contains all the frequencies (Figure 8.2d,e), but the Fourier transform of a sine wave has just one frequency. The smallest frequency is . = 0, for which the vector becomes x = (. . . , 1, 1, 1, 1, 1, . . .). The highest frequency is . = ., where the same vector becomes x = (. . . , 1,-1, 1,-1, 1, . . .). The special feature of this input is that the output vector y(n) is a multiple of the input. For the moving average, the output (filter response) is y(n) = 1 2x(n) + 1 2x(n - 1) = 1 2ein. + 1 2ei(n-1). = 1 2 + 1 2e-i. ein. = H(.)x(n), where H(.) = (1 2 + 1 2 e-i.) is the frequency response function of the filter. Since H(0) = 1/2 + 1/2 = 1, we see that the input x = (. . . , 1, 1, 1, 1, 1, . . .) is transformed to itself. Also, H(.) for small values of . generates output that is very similar to the input. This filter “lets” the low frequencies through, hence the name “lowpass filter.” For . = ., 8.7 Filter Banks 773 the input is x = (. . . , 1,-1, 1,-1, 1, . . .) and the output is all zeros (since the average of 1 and -1 is zero). This lowpass filter smooths out the high-frequency regions (the bumps) of the input signal. Notice that we can write H(.) = cos . 2 ei./2. When we plot the magnitude |H(.)| = cos(./2) of H(.) (Figure 8.27a), it is easy to see that it has a maximum at . = 0 (the lowest frequency) and two minima at . = ±. (the highest frequencies). The highpass filter uses differences to pick up the high frequencies in the input signal, and reduces or removes the smooth (low frequency) parts. In the case of the Haar transform, the highpass filter computes y(n) = 1 2x(n) - 1 2x(n - 1) = h  x, where the filter coefficients are h(0) = 1/2 and h(1) = -1/2, or h = (. . . , 0, 0, 1/2,-1/2, 0, . . .). In matrix notation this can be expressed by ..... · · · y(-1) y(0) y(1) · · · ..... =...... · · · -1 2 1 2 -1 2 1 2 -1 2 1 2 · · · ...... ..... · · · x(-1) x(0) x(1) · · · ..... . The main diagonal contains copies of h(0), and the diagonal below contains h(1). Using the identity and delay operator, this can also be written highpass filter = 1 2 (identity) - 1 2 (delay). Again selecting input x(n) = ein., it is easy to see that the output is y(n) = 1 2ein. - 1 2ei(n-1). = 1 2 - 1 2e-i. e-i./2 = sin (./2) ie-i./2. This time the highpass response function is H1(.) = 1 2 - 1 2e-i. = 1 2 ei./2 - e-i./2e-i./2 = sin (./2) e-i./2. The magnitude is |H1(.)| = | sin . 2 |. It is shown in Figure 8.27b, and it is obvious that it has a minimum for frequency zero and two maxima for large frequencies. 774 8. Wavelet Methods -. . -. . (a) (b) Figure 8.27: Magnitudes of (a) Lowpass and (b) Highpass Filters. An important property of filter banks is that none of the individual filters are invertible, but the bank as a whole has to be designed such that the input signal could be perfectly reconstructed from the output in spite of the data loss caused by downsampling. It is easy to see, for example, that the constant signal x = (. . . , 1, 1, 1, 1, 1, . . .) is transformed by the highpass filter H1 to an output vector of all zeros. Obviously, there cannot exist an inverse filter H-1 1 that will be able to reconstruct the original input from a zero vector. The best that such an inverse transform can do is to use the zero vector to reconstruct another zero vector.  Exercise 8.8: Show an example of an input vector x that is transformed by the lowpass filter H0 to a vector of all zeros. Summary: The discussion of filter banks in this section should be compared to the discussion of image transforms in Section 7.6. Even though both sections describe transforms, they differ in their approach, since they describe different classes of transforms. Each of the transforms described in Section 7.6 is based on a set of orthogonal basis functions (or orthogonal basis images), and is computed as an inner product of the input signal with the basis functions. The result is a set of transform coefficients that are subsequently compressed either losslessly (by RLE or some entropy encoder) or lossily (by quantization followed by entropy coding). This section deals with subband transforms [Simoncelli and Adelson 90], a different type of transform that is computed by taking the convolution of the input signal with a set of bandpass filters and decimating the results. Each decimated set of transform coefficients is a subband signal that encodes a specific range of the frequencies of the input. Reconstruction is done by upsampling, followed by computing the inverse transforms, and merging the resulting sets of outputs from the inverse filters. The main advantage of subband transforms is that they isolate the different frequencies of the input signal, thereby making it possible for the user to precisely control the loss of data in each frequency range. In practice, such a transform decomposes an image into several subbands, corresponding to different image frequencies, and each subband can be quantized differently. The main disadvantage of this type of transform is the introduction of artifacts, such as aliasing and ringing, into the reconstructed image, because of the downsampling. This is why the Haar transform is unsatisfactory, and most of the research in this field is concerned with finding better sets of filters. 8.7 Filter Banks 775 Figure 8.28 illustrates a typical case of a general subband filter bank with N bandpass filters and three stages. Notice how the output of the lowpass filter H0 of each stage is sent to the next stage for further decomposition, and how the combined output of the synthesis bank of a stage is sent to the top inverse filter of the synthesis bank of the preceding stage. H0 H1 vk0 vk1 HN vkN x(n) y0(n) y0(n) y0(n) y1(n) y1(n) y1(n) yN(n) yN(n) yN(n) x(n) H0 H1 vk0 vk1 HN vkN H0 H1 vk0 vk1 HN vkN F0 F1 ^k0 ^k1 ^kN FN F0 F1 ^k0 ^k1 ^kN FN F0 F1 ^k0 ^k1 ^kN FN stage 1 stage 2 stage 3 Figure 8.28: A General Filter Bank. 8.7.1 Deriving the Filter Coefficients After presenting the basic operation of filter banks, the natural question is, How are the filter coefficients derived? A full answer is outside the scope of this book (see, for example, [Akansu and Haddad 92]), but this section provides a glimpse at the rules and methods used to figure out the values of various filter banks. Given a set of two forward and two inverse N-tap filters H0 and H1, and F0 and F1 (where N is even), we denote their coefficients by h0 = h0(0), h0(1), . . . , h0(N - 1), h1 = h1(0), h1(1), . . . , h1(N - 1), f0 = f0(0), f0(1), . . . , f0(N - 1), f1 = f1(0), f1(1), . . . , f1(N - 1). The four vectors h0, h1, f0, and f1 are the impulse responses of the four filters. The simplest set of conditions that these quantities have to satisfy is: 1. Normalization: Vector h0 is normalized (i.e., its length is one unit). 2. Orthogonality: For any integer i that satisfies 1 . i < N/2, the vector formed by the first 2i elements of h0 should be orthogonal to the vector formed by the last 2i elements of the same h0. 3. Vector f0 is the reverse of h0. 776 8. Wavelet Methods 4. Vector h1 is a copy of f0 where the signs of the odd-numbered elements (the first, third, etc.) are reversed. We can express this by saying that h1 is computed by coordinate multiplication of h1 and (-1, 1,-1, 1, . . . ,-1, 1). 5. Vector f1 is a copy of h0 where the signs of the even-numbered elements (the second, fourth, etc.) are reversed. We can express this by saying that f1 is computed by coordinate multiplication of h0 and (1,-1, 1,-1, . . . , 1,-1). For two-tap filters, rule 1 implies h20 (0) + h20 (1) = 1. (8.9) Rule 2 is not applicable because N = 2, so i < N/2 implies i < 1. Rules 3–5 yield f0 = h0(1), h0(0), h1 = -h0(1), h0(0), f1 = h0(0),-h0(1). It all depends on the values of h0(0) and h0(1), but the single Equation (8.9) is not enough to determine them. However, it is not difficult to see that the choice h0(0) = h0(1) = 1/.2 satisfies Equation (8.9). For four-tap filters, rules 1 and 2 imply h20 (0) + h20 (1) + h20 (2) + h20 (3) = 1, h0(0)h0(2) + h0(1)h0(3) = 0, (8.10) and rules 3–5 yield f0 = h0(3), h0(2), h0(1), h0(0), h1 = -h0(3), h0(2),-h0(1), h0(0), f1 = h0(0),-h0(1), h0(2),-h0(3). Again, Equation (8.10) is not enough to determine four unknowns, and other considerations (plus mathematical intuition) are needed to derive the four values. They are listed in Equation (8.11) (this is the Daubechies D4 filter).  Exercise 8.9: Write the five conditions above for an eight-tap filter. Determining the N filter coefficients for each of the four filters H0, H1, F0, and F1 depends on h0(0) through h0(N - 1), so it requires N equations. However, in each of the cases above, rules 1 and 2 supply only N/2 equations. Other conditions have to be imposed and satisfied before the N quantities h0(0) through h0(N - 1) can be determined. Here are some examples: Lowpass H0 filter: We want H0 to be a lowpass filter, so it makes sense to require that the frequency response H0(.) be zero for the highest frequency . = .. Minimum phase filter: This condition requires the zeros of the complex function H0(z) to lie on or inside the unit circle in the complex plane. Controlled collinearity: The linearity of the phase response can be controlled by requiring that the sum i h0(i) - h0(N - 1 - i)2 be a minimum. Other conditions are discussed in [Akansu and Haddad 92]. 8.8 The DWT 777 8.8 The DWT Information that is produced and analyzed in real-life situations is discrete. It comes in the form of numbers, rather than a continuous function. This is why the discrete rather than the continuous wavelet transform is the one used in practice ([Daubechies 88], [De- Vore et al. 92], and [Vetterli and Kovacevic 95]). Recall that the CWT [Equation (8.3)] is the integral of the product f(t).*( t-b a ), where a, the scale factor, and b, the time shift, can be any real numbers. The corresponding calculation for the discrete case (the DWT) involves a convolution, but experience shows that the quality of this type of transform depends heavily on two factors, the choice of scale factors and time shifts, and the choice of wavelet. In practice, the DWT is computed with scale factors that are negative powers of 2 and time shifts that are nonnegative powers of 2. Figure 8.29 shows the so-called dyadic lattice that illustrates this particular choice. The wavelets used are those that generate orthonormal (or biorthogonal) wavelet bases. The main thrust in wavelet research has therefore been the search for wavelet families that form orthogonal bases. Of those wavelets, the preferred ones are those that have compact support, because they allow for DWT computations with finite impulse response (FIR) filters. Scale Time 1 20 21 22 23 2 3 Figure 8.29: The Dyadic Lattice Showing the Relation Between Scale Factors and Time. The simplest way to describe the discrete wavelet transform is by means of matrix multiplication, along the lines developed in Section 8.6.3. The Haar transform depends on two filter coefficients c0 and c1, both with a value of 1/.2 . 0.7071. The smallest transform matrix that can be constructed in this case is 1 1 1-1/.2. It is a 2Χ2 matrix, and it generates two transform coefficients, an average and a difference. (Notice that these are not exactly an average and a difference, because .2 is used instead of 2. Better names for them are coarse detail and fine detail, respectively.) In general, the DWT can 778 8. Wavelet Methods use any set of wavelet filters, but it is computed in the same way regardless of the particular filter used. We start with one of the most popular wavelets, the Daubechies D4. As its name implies, it is based on four filter coefficients c0, c1, c2, and c3, whose values are listed in Equation (8.11). The transform matrix W is [compare with matrix A1, Equation (8.6)] W = .............. c0 c1 c2 c3 0 0 . . . 0 c3 -c2 c1 -c0 0 0 . . . 0 0 0 c0 c1 c2 c3 . . . 0 0 0 c3 -c2 c1 -c0 . . . 0 ... ... . . . 0 0 . . . 0 c0 c1 c2 c3 0 0 . . . 0 c3 -c2 c1 -c0 c2 c3 0 . . . 0 0 c0 c1 c1 -c0 0 . . . 0 0 c3 -c2 .............. . When this matrix is applied to a column vector of data items (x1, x2, . . . , xn), its top row generates the weighted sum s1 = c0x1 + c1x2 + c2x3 + c3x4, its third row generates the weighted sum s2 = c0x3 + c1x4 + c2x5 + c3x6, and the other odd-numbered rows generate similar weighted sums si. Such sums are convolutions of the data vector xi with the four filter coefficients. In the language of wavelets, each of them is called a smooth coefficient, and together they are called an H smoothing filter. In a similar way, the second row of the matrix generates the quantity d1 = c3x1 - c2x2 + c1x3 - c0x4, and the other even-numbered rows generate similar convolutions. Each di is called a detail coefficient, and together they are called a G filter. G is not a smoothing filter. In fact, the filter coefficients are chosen such that the G filter generates small values when the data items xi are correlated. Together, H and G are called quadrature mirror filters (QMF). The discrete wavelet transform of an image can therefore be viewed as passing the original image through a QMF that consists of a pair of lowpass (H) and highpass (G) filters. If W is an nΧn matrix, it generates n/2 smooth coefficients si and n/2 detail coefficients di. The transposed matrix is WT = .............. c0 c3 0 0 . . . c2 c1 c1 -c2 0 0 . . . c3 -c0 c2 c1 c0 c3 . . . 0 0 c3 -c0 c1 -c2 . . . 0 0 . . . c2 c1 c0 c3 0 0 c3 -c0 c1 -c2 0 0 c2 c1 c0 c3 c3 -c0 c1 -c2 .............. . It can be shown that in order for W to be orthonormal, the four coefficients have to satisfy the two relations c20 + c21 + c22 + c23 = 1 and c2c0 + c3c1 = 0. The other two 8.8 The DWT 779 equations used to calculate the four filter coefficients are c3 - c2 + c1 - c0 = 0 and 0c3 - 1c2 +2c1 - 3c0 = 0. They represent the vanishing of the first two moments of the sequence (c3,-c2, c1,-c0). The solutions are c0 = (1+.3)/(4.2) . 0.48296, c1 = (3+.3)/(4.2) . 0.8365, c2 = (3 - .3)/(4.2) . 0.2241, c3 = (1 - .3)/(4.2) . -0.1294. (8.11) Using a transform matrix W is conceptually simple, but not very practical, since W should be of the same size as the image, which can be large. However, a look at W shows that it is very regular, so there is really no need to construct the full matrix. It is enough to have just the top row of W. In fact, it is enough to have just an array with the filter coefficients. Figure 8.30 lists Matlab code that performs this calculation. Function fwt1(dat,coarse,filter) takes a row vector dat of 2n data items, and another array, filter, with filter coefficients. It then calculates the first coarse levels of the discrete wavelet transform.  Exercise 8.10: Write similar code for the inverse one-dimensional discrete wavelet transform. Plotting Functions: Wavelets are being used in many fields and have many applications, but the simple test of Figure 8.30 suggests another application, namely, plotting functions. Any graphics program or graphics software package has to include a routine to plot functions. It works by calculating the function at certain points and connecting the points with straight segments. In regions where the function has small curvature (it resembles a straight line) only few points are needed, whereas in areas where the function has large curvature (it changes direction rapidly) more points are required. An ideal plotting routine should therefore be adaptive. It should select the points depending on the curvature of the function. The curvature, however, may not be easy to compute (it is essentially given by the second derivative of the function) which is why many plotting routines use instead the angle between consecutive segments. Figure 8.31 shows how a typical plotting routine works. It starts with a fixed number (say, 50) of points. This implies 49 straight segments connecting them. Before any of the segments is actually plotted, the routine measures the angles between consecutive segments. If an angle at point Pi is extreme (close to zero or close to 360., as it is around points 4 and 10 in the figure), then more points are calculated between points Pi-1 and Pi+1; otherwise (if the angle is closer to 180., as, for example, around points 5 and 9 in the figure), Pi is considered the only point necessary in that region. Better and faster results may be obtained using a discrete wavelet transform. The function is evaluated at n points (where n, a parameter, is large), and the values are collected in a vector v. A discrete wavelet transform of v is then calculated, to produce n transform coefficients. The next step is to discard m of the smallest coefficients (where m is another parameter). We know, from the previous discussion, that the smallest coefficients represent small details of the function, so discarding them leaves the important details practically untouched. The inverse transform is then performed on the remaining n - m transform coefficients, resulting in n - m new points that are 780 8. Wavelet Methods function wc1=fwt1(dat,coarse,filter) % The 1D Forward Wavelet Transform % dat must be a 1D row vector of size 2^n, % coarse is the coarsest level of the transform % (note that coarse should be < 1 is a real number, w is written either as w = bh, with 0 < h < 1, or as w = bd-2, with 1 < d < 2, and i = .-1. Notice that Wb,w(t) is complex; its real and imaginary parts are called the Weierstrass cosine and sine functions, respectively. Weierstrass proved the unusual behavior of this function, and also showed that for d < 1 it is differentiable. In his time, such a function was so contrary to common sense and mathematical intuition that he did not publish his findings. Today, we simply call this function and others like it fractals.  Exercise 8.11: Use functions fwt2 and iwt2 of Figures 8.33 and 8.34 to blur an image. The idea is to compute the 4-step subband transform of an image (thus ending up with 13 subbands), then set most of the transform coefficients to zero and heavily quantize some of the others. This, of course, results in a loss of image information, and in a nonperfectly reconstructed image. The aim of this exercise, however, is to have the inverse transform produce a blurred image. This illustrates an important property of the discrete wavelet transform, namely its ability to reconstruct images that degrade gracefully when more and more transform coefficients are zeroed or coarsely quantized. Other transforms, most notably the DCT, may introduce artifacts in the reconstructed image, but this property of the DWT makes it ideal for applications such as fingerprint compression (Section 8.18). 790 8. Wavelet Methods 8.9 Multiresolution Decomposition The main idea of wavelet analysis, illustrated in detail in Section 8.5, is to analyze a function at different scales. A mother wavelet is used to construct wavelets at different scales (dilations) and translate each relative to the function being analyzed. The results of translating a wavelet depend on how much it matches the function being analyzed. Wavelets at different scales (resolutions) produce different results. The principle of multiresolution decomposition, due to Stephane Mallat and Yves Meyer, is to group all the wavelet transform coefficients for a given scale, display their superposition, and repeat for all scales. Figure 8.39 lists a Matlab function multres for this operation, together with a test. Figure 8.38 shows two examples, a single spike and multiple spikes. It is easy to see how the fine-resolution coefficients are concentrated at the values of t that correspond to the spikes (i.e., the high activity areas of the data). 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 0 500 1000 1500 2000 2500 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 -10 -9 -8 -7 -6 -5 -4 -3 -2 -1 0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 -10 -9 -8 -7 -6 -5 -4 -3 -2 -1 0 (a) (b) Figure 8.38: Examples of Multiresolution Decomposition. (a) A Spike. (b) Several Spikes. 8.10 Various Image Decompositions 791 function multres(wc,coarse,filter) % A multi resolution plot of a 1D wavelet transform scale=1./max(abs(wc)); n=length(wc); j=log2(n); LockAxes([0 1 -(j) (-coarse+2)]); t=(.5:(n-.5))/n; for i=(j-1):-1:coarse z=zeros(1,n); z((2^(i)+1):(2^(i+1)))=wc((2^(i)+1):(2^(i+1))); dat=iwt1(z,i,filter); plot(t,-(i)+scale.*dat); end z=zeros(1,n); z(1:2^(coarse))=wc(1:2^(coarse)); dat=iwt1(z,coarse,filter); plot(t,(-(coarse-1))+scale.*dat); UnlockAxes; And a test routine n=1024; t=(1:n)./n; dat=spikes(t); % several spikes %p=floor(n*.37); dat=1./abs(t-(p+.5)/n); % one spike figure(1), plot(t,dat) filt=[0.4830 0.8365 0.2241 -0.1294]; wc=fwt1(dat,2,filt); figure(2), plot(t,wc) figure(3) multres(wc,2,filt); function dat=spikes(t) pos=[.05 .11 .14 .22 .26 .41 .44 .64 .77 .79 .82]; hgt=[5 5 4 4.5 5 3.9 3.3 4.6 1.1 5 4]; wth=[.005 .005 .006 .01 .01 .03 .01 .01 .005 .008 .005]; dat=zeros(size(t)); for i=1:length(pos) dat=dat+hgt(i)./(1+abs((t-pos(i))./wth(i))).^4; end; Figure 8.39: Matlab Code for the Multiresolution Decomposition of a One-Dimensional Row Vector. 8.10 Various Image Decompositions Section 8.6.1 shows two ways of applying a discrete wavelet transform to an image in order to partition it into several subbands. This section (based on [Strψmme 99], which also contains comparisons of experimental results) discusses seven ways of doing the same thing, each involving a different algorithm and resulting in subbands with different energy compactions. Other decompositions are also possible (Section 8.18 describes a special, symmetric decomposition). It is important to realize that the wavelet filters and the decomposition method are independent. The discrete wavelet transform of an image can use any set of wavelet 792 8. Wavelet Methods filters and can decompose the image in any way. The only limitation is that there must be enough data points in the subband to cover all the filter taps. For example, if a 12-tap Daubechies filter is used, and the image and subband sizes are powers of two, then the smallest subband that can be produced has size 8Χ8. This is because a subband of size 16Χ16 is the smallest that can be multiplied by the 12 coefficients of this particular filter. Once such a subband is decomposed, the resulting 8Χ8 subbands are too small to be multiplied by 12 coefficients and to be decomposed further. 1. Laplacian Pyramid: This technique for image decomposition is described in detail in Section 8.13. Its main feature is progressive image transmission. During decompression and image reconstruction, the user sees small, blurred images that grow and become sharper. The main reference is [Burt and Adelson 83]. The Laplacian pyramid is generated by subtracting an upsampled lowpass version of the image from the original image. The image is partitioned into a Gaussian pyramid (the lowpass subbands) and a Laplacian pyramid that consists of the detail coefficients (the highpass subbands). Only the Laplacian pyramid is needed to reconstruct the image. The transformed image is bigger than the original image, which is the main difference between the Laplacian pyramid decomposition and the pyramid decomposition (method 4 below). 2. Line: This technique is a simpler version of the standard wavelet decomposition (method 5). The wavelet transform is applied to each row of the image, resulting in smooth coefficients on the left (subband L1) and detail coefficients on the right (subband H1). Subband L1 is then partitioned into L2 and H2, and the process is repeated until the entire coefficient matrix is turned into detail coefficients, except the leftmost column, which contains smooth coefficients. The wavelet transform is then applied recursively to the leftmost column, resulting in one smooth coefficient at the top-left corner of the coefficient matrix. This last step may be omitted if the compression method being used requires that image rows be individually compressed (notice the distinction between the wavelet transform and the actual compression algorithm). This technique exploits correlations only within an image row to calculate the transform coefficients. Also, discarding a coefficient that is located on the leftmost column may affect just a particular group of rows and may this way introduce artifacts into the reconstructed image. Implementation of this method is simple, and execution is fast, about twice that of the standard decomposition. This type of decomposition is illustrated in Figure 8.40. It is possible to apply this decomposition to the columns of the image, instead of to the rows. Ideally, the transform should be applied in the direction of highest image redundancy, and experience suggests that for natural images this is the horizontal direction. Thus, in practice, line decomposition is applied to the image rows. 3. Quincunx: Somewhat similar to the Laplacian pyramid, quincunx decomposition proceeds level by level and decomposes subband Li of level i into subbands Hi+1 and Li+1 of level i + 1. Figure 8.41 illustrates this type of decomposition. The method is due to Strψmme and McGregor [Strψmme and McGregor 97], who originally called it nonstandard decomposition. (See also [Starck et al. 98] for a different presentation of this method.) It is efficient and computationally simple. On average, it achieves more than four times the energy compaction of the line method. 8.10 Various Image Decompositions 793 L3H1 L3H1 L3H2 L3L2 L3L1 H3 H2 H1 H3 H2 H1 L3H3 H2 H1 L1 L2 H2 Original image H1 H1 Figure 8.40: Line Wavelet Decomposition. Quincunx decomposition results in fewer subbands than most other wavelet decompositions, a feature that may lead to reconstructed images with slightly lower visual quality. The method is not used much in practice, but [Strψmme 99] presents results that suggest that the quincunx decomposition performs extremely well and may be the best performer in many practical situations. 4. Pyramid: The pyramid decomposition is by far the most common method used to decompose images that are wavelet transformed. It results in subbands with horizontal, vertical, and diagonal image details, as illustrated by Figure 8.17. The three subbands at each level contain horizontal, vertical, and diagonal image features at a particular scale, and each scale is divided by an octave in spatial frequency (division of the frequency by two). Pyramid decomposition turns out to be a very efficient way of transferring significant visual data to the detail coefficients. Its computational complexity is about 30% higher than that of the quincunx method, but its image reconstruction abilities are higher. The reasons for the popularity of the pyramid method may be that (1) it is symmetrical, (2) its mathematical description is simple, and (3) it was used by the influential paper [Mallat 89]. Figure 8.42 illustrates pyramid decomposition. It is obvious that the first step is identical to that of the quincunx decomposition. However, while the quincunx method leaves the high-frequency subband untouched, the pyramid method resolves it into two bands. On the other hand, pyramid decomposition involves more computations in order to spatially resolve the asymmetric high-frequency band into two symmetric highfrequency and low-frequency bands. 5. Standard: The first step in the standard decomposition is to apply whatever discrete wavelet filter is being used to all the rows of the image, obtaining subbands L1 and H1. This is repeated on L1 to obtain L2 and H2, and so on k times. This is followed by a second step where a similar calculation is applied k times to the columns. If k = 1, 794 8. Wavelet Methods L1 H1 L2 H1 H2 L3 H1 H3 H2 L4 H1 H3 H2 H4 Original image Figure 8.41: Quincunx Wavelet Decomposition. From the Dictionary QUINCUNX: An arrangement of five objects in a square or rectangle, one at each corner, and one at the center. The word seems to have originated from Latin around 1640– 1650. It stands for “five-twelvths.” Quinc is a variation of quinque, and unx or unc is a form of uncia, meaning twelfth. It was used to indicate a Roman coin worth five-twelfths of an AS and marked with a quincunx of spots. (AS is an ancient Roman unit of weight, equal to about 12 ounces.)      temporary L Original temporary L temporary H temporary H LL1 LH1 HL1 HH1 LL2 LH2 HL2 HH2 image Figure 8.42: Pyramid Wavelet Decomposition. 8.10 Various Image Decompositions 795 the decomposition alternates between rows and columns, but k may be greater than 1. The end result is to have one smooth coefficient at the top-left corner of the coefficient matrix. This method is somewhat similar to line decomposition. An important feature of standard decomposition is that when a coefficient is quantized, it may affect a long, thin rectangular area in the reconstructed image. Thus, very coarse quantization may result in artifacts in the reconstructed image in the form of horizontal rectangles. Standard decomposition has the second-highest reconstruction quality of the seven methods described here. The reason for the improvement compared to the pyramid decomposition may be that the higher directional resolution gives thresholding a better chance to cover larger uniform areas. On the other hand, standard decomposition is computationally more expensive than pyramid decomposition. Original image L1 H1 L2 H2 H1 H2 H1 L3 H3 Figure 8.43: Standard Wavelet Decomposition. 6. Uniform Decomposition: This method is also called the wavelet packet transform. It is illustrated in Figure 8.44. In the case where the user elects to compute all the levels of the transform, the uniform decomposition becomes similar to the discrete Fourier transform (DFT) in that each coefficient represents a spatial frequency for the entire image. In such a case (where all the levels are computed), the removal of one coefficient in the transformed image affects the entire reconstructed image. The computational cost of uniform decomposition is very high, since it effectively computes n2 coefficients for every level of decomposition, where n is the side length of the (square) image. Despite having comparably high average reconstruction qualities, the perceptual quality of the image starts degrading at lower ratios than for the other decomposition methods. The reason for this is the same as for Fourier methods: Because the support for a single coefficient is global, its removal has the effect of blurring the reconstructed image. The conclusion is that the increased computational complexity of uniform decomposition does not result in increased quality of the reconstructed image. 6.1. Full Wavelet Decomposition: (This is a special case of uniform decomposition.) Denote the original image by I0. We assume that its size is 2lΧ2l. After applying the discrete wavelet transform to it, we end up with a matrix I1 partitioned into four subbands. The same discrete wavelet transform (i.e., using the same wavelet filters) is then applied recursively to each of the four subbands individually. The result 796 8. Wavelet Methods LHLH LHHH HHLH HHHH LHLL LHHL HHLL HHHL LLLH LLHH HLLH HLHH LLLL LLHL HLLL HLHL Original Temporary L image Temporary H Temporary L Temporary H Temporary L Temporary H Temporary L Temporary H Temporary L Temporary H LL LH HL HH LLLL LLLH LLHL LLHH LHLL LHLH LHHL LHHH HLLL HLLH HLHL HLHH HHLL HHLH HHHL HHHH Figure 8.44: Uniform Wavelet Decomposition. is a coefficient matrix I2 consisting of 16 subbands. When this process is carried out r times, the result is a coefficient matrix consisting of 2rΧ2r subbands, each of size 2l-rΧ2l-r. The top-left subband contains the smooth coefficients (depending on the particular wavelet filter used, it may look like a small version of the original image), and the other subbands contain detail coefficients. Each subband corresponds to a frequency band, while each individual transform coefficient corresponds to a local spatial region. By increasing the recursion depth r, we can increase frequency resolution at the expense of spatial resolution. This type of wavelet image decomposition has been proposed by Wong and Kuo [Wong and Kuo 93], who highly recommend its use. However, it seems that it has been ignored by researchers and implementers in the field of image compression. 7. Adaptive Wavelet Packet Decomposition: The uniform decomposition method is costly in terms of computations, and the adaptive wavelet packet method is potentially even more so. The idea is to skip those subband splits that do not contribute significantly to energy compaction. The result is a coefficient matrix with subbands of different (possibly even many) sizes. The bottom-right part of Figure 8.45 shows such a case (after [Meyer et al. 98], which shows the adaptive wavelet packet transform matrix for the bi-level “mandril” image, Figure 7.56). The justification for this complex decomposition method is the prevalence of continuous tone (natural) images. These images, which are discussed at the start of Chapter 7, are mostly smooth but normally also have some regions with high frequency data. Such 8.10 Various Image Decompositions 797 regions should end up as many small subbands (to better enable an accurate spatial frequency representation of the image), with the rest of the image giving rise to a few large subbands. The bottom-left coefficient matrix in Figure 8.45 is an example of a very uniform image, resulting in just 10 subbands. (The test for a split depends on the absolute magnitude of the transform coefficients. Thus, the test can be adjusted so high that very few splits are done.) LL LLLL LLLH LLHL LLHH Temporary L Temporary H Temporary L Original image LH HL HLLL HLLH HLHL HLHH HH Temporary L Temporary H Temporary H LH HH LLLL LLLH LLHL LLHH HLLL HLLH HLHL HLHH Figure 8.45: Adaptive Wavelet Packet Decomposition. The downside of this type of decomposition is finding an algorithm that will determine which subband splits can be skipped. Such an algorithm uses entropy calculations and should be efficient. It should identify all the splits that do not have to be performed, and it should identify as many of them as possible. An inefficient algorithm may lead to the split of every subband, thereby performing many unnecessary computations and ending up with a coefficient matrix where every coefficient is a subband, in which case this decomposition reduces to the uniform decomposition. 798 8. Wavelet Methods This type of decomposition has the highest reproduction quality of all the methods discussed here, a feature that may justify the high computational costs in certain special applications. This quality, however, is not much higher than what is achieved with simpler decomposition methods, such as standard, pyramid, or quincunx. 8.11 The Lifting Scheme The lifting scheme ([Stollnitz et al. 96] and [Sweldens and Schr¨oder 96]) is a novel and useful way of looking at the discrete wavelet transform. It is easy to understand, since it performs all the operations in the time domain, rather than in the frequency domain, and has other advantages as well. This section illustrates the lifting approach using the Haar transform, which is already familiar to the reader, as an example. The following section extends the same approach to other transforms. The Haar transform, described Section 8.6, is based on the computations of averages and differences (details). Given two adjacent pixels a and b, the principle is to calculate the average s = (a + b)/2 and difference d = b - a. If a and b are similar, s will be similar to both and d will be small, i.e., require fewer bits to represent. This transform is reversible, because a = s - d/2 and b = s + d/2, and it can be written using matrix notation as (s, d) = (a, b) 1/2 -1 1/2 1 = (a, b)A, (a, b) = (s, d) 1 1 -1/2 1/2 = (s, d)A-1. Consider a row of 2n pixel values sn,l for 0 . l < 2n. There are 2n-1 pairs of pixels sn,2l, sn,2l+1 for l = 0, 2, 4, . . . , 2n-2. Each pair is transformed into an average sn-1,l = (sn,2l + sn,2l+1)/2 and a difference dn-1,l = sn,2l+1 - sn,2l. The result is a set sn-1 of 2n-1 averages and a set dn-1 of 2n-1 differences. The same operations can be applied to the 2n-1 averages sn-1,l of set sn-1, resulting in 2n-2 averages sn-2,l and 2n-2 differences dn-2,l. After applying these operations n times we end up with a set s0 consisting of one average s0,0, and with n sets of differences dj,l where j = 0, 1, . . . , n - 1 and l = 0, 1, . . . , 2j - 1. Set j consists of j differences, so the total number of differences is n-1 j=0 2j = 2n - 1. Adding the single average s0,0 brings the total number of results to 2n, the same as the number of original pixel values. Notice that the final average s0,0 is the average S of the original 2n pixel values, so it can be called the DC component of the original values. In fact, if we look at any set sj of averages sj,l, l = 0, 1, . . . , 2j -1, we find that its average S = 1 2j 2j-1 l=0 sj,l 8.11 The Lifting Scheme 799 dn-1 dn-2 d1 d0 d0 d1 dn-2 dn-1 sn sn-1 sn-2 s1 s0 s0 s1 s2 sn-1 sn (a) (b) Figure 8.46: (a) The Haar Wavelet Transform and (b) Its Inverse. is the average of all the original 2n pixel values. Thus, the average S of set sj is independent of j. Figure 8.46a,b illustrates the transform and its inverse. The main idea in the lifting scheme is to perform all the necessary operations without using extra space. The entire transform is performed in place, replacing the original image. We start with a pair of consecutive pixels a and b. They are replaced with their average s and difference d by first replacing b with d = b - a, then replacing a with s = a + d/2 [since d = b - a, a + d/2 = a + (b - a)/2 equals (a + b)/2]. In the C language, this is written b-=a; a+=b/2;  Exercise 8.12: Write the reverse operations in C. This is easy to apply to an entire row of pixels. Suppose that we have a row sj with 2j values and we want to transform it to a row sj-1 with 2j-1 averages and 2j-1 differences. The lifting scheme performs this transform in three steps, split, predict, and update. The split operation splits row sj into two separate sets denoted by evenj-1 and oddj-1. The former contains all the even values sj,2l, and the latter contains all the odd values sj,2l+1. The range of l is from 0 to 2j - 1. This kind of splitting into odd and even values is called the lazy wavelet transform. We denote it by (evenj-1, oddj-1) := Split(sj). The predict operation uses the even set evenj-1 to predict the odd set oddj-1. This is based on the fact that each value sj,2l+1 in the odd set is adjacent to the corresponding value sj,2l in the even set. Thus, the two values are correlated and either can be used to predict the other. Recall that a general difference dj-1,l is computed as the difference dj-1,l = sj,2l+1 -sj,2l between an odd value and an adjacent even value (or between an odd value and its prediction), which is why we can define the prediction operator P as dj-1 = oddj-1 - P(evenj-1). 800 8. Wavelet Methods The update operation U follows the prediction step. It calculates the 2j-1 averages sj-1,l as the sum sj-1,l = sj,2l + dj-1,l/2. (8.13) This operation is formally defined by sj-1 = evenj-1 + U(dj-1).  Exercise 8.13: Use Equation (8.13) to show that sets sj and sj-1 have the same average. The important point to notice is that all three operations can be performed in place. The even locations of set sj are overwritten with the averages (i.e., with set evenj-1), and the odd locations are overwritten with the differences (set oddj-1). The sequence of three operations can be summarized by (oddj-1, evenj-1) := Split(sj ); oddj-1- = P(evenj-1); evenj-1+ = U(oddj-1); Figure 8.47a is a wiring diagram of this process. (b) sj-1 dj-1 oddj-1 evenj-1 U P Merge sj + - (a) sj-1 dj-1 oddj-1 evenj-1 sj Split P U + - Figure 8.47: The Lifting Scheme. (a) Forward Transform. (b) Reverse Transform. The reverse transform is similar. It is based on the three operations undo update, undo prediction, and merge. Given two sets sj-1 and dj-1, the “undo update” operation reconstructs the averages sj,2l (the even values of set sj) by subtracting the update operation. It generates the set evenj-1 by subtracting the sets sj-1-U(dj-1). Written explicitly, this operation becomes sj,2l = sj-1,l - dj-1,l/2, for 0 . l < 2j . 8.11 The Lifting Scheme 801 Given the two sets evenj-1 and dj-1, the “undo predict” operation reconstructs the differences sj,2l+1 (the odd values of set sj) by adding the prediction operator P. It generates the set oddj-1 by adding the sets dj-1 + P(evenj-1). Written explicitly, this operation becomes sj,2l+1 = dj-1,l + sj,2l, for 0 . l < 2j . Now that the two sets evenj-1 and oddj-1 have been reconstructed, they are merged by the “merge” operation into the set sj . This is the inverse lazy wavelet transform, formally denoted by sj = Merge(evenj-1, oddj-1). It moves the averages and the differences into the even and odd locations of sj , respectively, without using any extra space. The three operations are summarized by evenj-1- = U(oddj-1); oddj-1+ = P(evenj-1); sj := Merge(oddj-1, evenj-1); Figure 8.47b is a wiring diagram of this process. The wiring diagrams show one of the important features of the lifting scheme, namely its inherent parallelism. Given an SIMD (single instruction, multiple data) computer with 2n processing units, each unit can be “responsible” for one pixel value. Such a computer executes a single program where each instruction is executed in parallel by all the processing units. Another advantage of lifting is the simplicity of its inverse transform. It is simply the code for the forward transform, run backward. The main advantage of lifting is the fact that it is easy to extend. It has been presented here for the Haar transform, where averages and differences are calculated in a simple way. It can be extended to more complex cases, where the prediction and update operations are more complex. 8.11.1 The Linear Wavelet Transform The reason for extending the lifting scheme beyond the Haar transform is that this transform does not produce high-quality results, since it uses such simple prediction. Recall that the “predict” operation uses the even set evenj-1 to predict the odd set oddj-1. This gives accurate prediction only in the (very rare) cases where the two sets are identical. We can say that the Haar transform eliminates order-zero correlation between pixels by using an order-one predictor. The Haar transform also preserves the average of all the pixels of the image, and this average can be called the order-zero moment of the image. Better compression can be achieved by transforms that use better predictors, predictors that exploit correlations between several neighbor pixels and also preserve higherorder moments of the image. The predictor and update operations described here are of order two. This implies that the predictor will provide exact prediction if the image pixels vary linearly, and the update operation will preserve the first two (order-zero and order-one) moments. The principle is easy to describe. We again concentrate on a row sj of pixel values. An odd-numbered value sj,2l+1 is predicted as the average of its two immediate neighbors sj,2l and sj,2l+2. To be able to reconstruct sj,2l+1, we have to calculate a detail value dj,l that is no longer a simple difference but is given by dj,l = sj,2l+1 - 1 2 (sj,2l + sj,2l+2). 802 8. Wavelet Methods (This notation assumes that every pixel has two immediate neighbors. This is not true for pixels located on the boundaries of the image, so edge rules will have to be developed and justified.) Figure 8.48 illustrates the meaning of the detail values in this case. Figure 8.48a shows the values of nine hypothetical pixels numbered 0 through 8. Figure 8.48b shows straight segments connecting the even-numbered pixels. This is the reason for the name “linear transform.” In Figure 8.48c the odd-numbered pixels are predicted by these straight segments, and Figure 8.48d shows (in dashed bold) the difference between each odd-numbered pixel and its prediction. The equation of a straight line is y = ax + b, a polynomial of degree 1, which is why we can think of the detail values as the amount by which the pixel values deviate locally from a degree-1 polynomial. If pixel x had value ax+b, all the detail values would be zero. This is why we can say that the detail values capture the high frequencies of the image. (a) (b) (c) (d) 0 1 2 3 4 5 6 7 8 Figure 8.48: Linear Prediction. The “update” operation reconstructs the averages sj-1,l from the averages sj,l and the differences dj-1,l. In the case of the Haar transform, this operation is defined by Equation (8.13): sj-1,l = sj,2l + dj-1,l/2. In the case of the linear transform the operation is somewhat more complicated. We derive the update operation for the linear transform using the requirement that it preserves the zeroth-order moment of the image, i.e., the average of the averages sj,l should not depend on j. We try an update of the form sj-1,l = sj,2l + A(dj-1,l-1 + dj-1,l), (8.14) 8.11 The Lifting Scheme 803 where A is an unknown coefficient. The sum of sj-1,l now becomes l sj-1,l = l sj,2l + 2A l dj-1,l = (1 - 2A) l sj,2l + 2A l sj,2l+1, so the choice A = 1/4 results in 2j-1-1 l=0 sj-1,l = 1 - 1 2 2j-1 l=0 sj,2l + 2 4 2j-1 l=0 sj,2l+1 = 1 2 2j-1 l=0 sj,l. Comparing this with Equation (8.15) shows that the zeroth-order moment of the image is preserved by the update operation of Equation (8.14). A direct check also verifies that l l sj-1,l = 1 2 l l sj,l, which shows that this update operation also preserves the first-order moment of the image and is therefore of order 2. Equation (8.15), duplicated below, is derived in the answer to Exercise 8.13: 2j-1-1 l=0 sj-1,l = 2j-1-1 l=0 (sj,2l + dj-1,l/2) = 1 2 2j-1-1 l=0 (sj,2l + sj,2l+1) = 1 2 2j-1 l=0 sj,l. (8.15) The inverse linear transform reconstructs the even and odd average values by sj,2l = sj-1,l - 1 4 (dj-1,l-1 + dj-1,l), and sj,2l+1 = dj,l + 1 2 (sj,2l + sj,2l+2), respectively. Figure 8.49 summarizes the linear transform. The top row shows the even and odd averages sj,l. The middle row shows how a detail value dj-1,l is calculated as the difference between an odd average sj,2l+1 and half the sum of its two even neighbors sj,2l and sj,2l+2. The bottom row shows how the next set sj-1 of averages is calculated. Each average sj-1,l in this set is the sum of an even average sj,2l and one-quarter of the two detail values dj-1,l-1 and dj-1,l. The figure also illustrates the main feature of the lifting scheme, namely how the even averages sj,2l are replaced by the next set of averages sj-1,l and how the odd averages sj,2l+1 are replaced by the detail values dj-1,l (the dashed arrows indicate items that are not moved). All these operations are performed in place. 8.11.2 Interpolating Subdivision The method of interpolating subdivision starts with a set s0 of pixels where pixel s0,k (for k = 0, 1, . . .) is stored in location k of an array a. It creates a set s1 (twice the size of s0) of pixels s1,k such that the even-numbered pixels s1,2k are simply the even-numbered 804 8. Wavelet Methods sj-1,l sj-1,l+1 sj,2l sj,2l sj,2l+1 sj,2l+2 sj,2l+2 dj-1,l dj-1,l dj-1,l-1 dj-1,l+1 -1/2 -1/2 -1/2 1/4 1/4 -1/2 1/4 1/4 Figure 8.49: Summary of the Linear Wavelet Transform. pixels s0,k and each of the odd-numbered pixels s1,2k+1 is obtained by interpolating some of the pixels (odd and/or even) of set s0. The new content of array a is s1,0 = s0,0, s1,1, s1,2 = s0,1, s1,3, s1,4 = s0,2, s1,5, . . . , s1,2k = s0,k, s1,2k+1, . . . . The original elements s0,k of s0 are now stored in a in locations 2k = k·21. We say that set s1 was created from s0 by a process of subdivision (or refinement). Next, set s2 (twice the size of s1) is created in the same way from s1. The evennumbered pixels s2,2k are simply the even-numbered pixels s1,k, and each of the oddnumbered pixels s2,2k+1 is obtained by interpolating some of the pixels (odd and/or even) of set s1. The new content of array a becomes s2,0 = s1,0 = s0,0, s2,1, s2,2 = s1,1, s2,3, s2,4 = s1,2 = s0,1, s2,5, . . . s2,2k = s1,k, s2,2k+1, . . . , s2,4k = s1,2k = s0,k, s2,4k+1, . . . . The original elements s0,k of s0 are now stored in a in locations 4k = k·22. In a general subdivision step, a set sj of pixel values sj,k is used to construct a new set sj+1, twice as large. The even-numbered pixels sj+1,2k are simply the even-numbered pixels sj,k, and each of the odd-numbered pixels sj+1,2k+1 is obtained by interpolating some of the pixels (odd and/or even) of set sj . The original elements s0,k of s0 are now stored in a in locations k·2j . We employ linear interpolation to illustrate this refinement process. Each of the odd-numbered pixels s1,2k+1 is calculated as the average of the two pixels s0,k and s0,k+1 of set s0. In general, we get sj+1,2k = sj,k, sj+1,2k+1 = 1 2 (sj,k + sj,k+1) . (8.16) Figure 8.50a–d shows several steps in this process, starting with a set s0 of four pixels (it is useful to visualize each pixel as a two-dimensional point). It is obvious that the pixel values converge to the polyline that passes through the original four points. Given an image where the pixel values go up and down linearly (in a polyline), we can compress it by selecting the pixels at the corners of the polyline (i.e., those pixels 8.11 The Lifting Scheme 805 (a) (b) (c) (d) Χ Χ Χ 0 0 0 0 Χ Χ Χ Χ Χ Χ 0 0 0 0 0 0 0 Χ Χ Χ Χ Χ Χ Χ Χ Χ Χ Χ Χ 0 0 0 0 0 0 0 0 0 0 0 0 0 Figure 8.50: An Example of a Linear Subdivision. where the value changes direction) and writing them on the compressed file. The image could then be perfectly reconstructed to any resolution by using linear subdivision to compute as many pixels as needed between corner pixels. Most images, however, feature more complex behavior of pixel values, so more complex interpolation is needed. We now show how to extend linear subdivision to polynomial subdivision. Instead of computing an odd-numbered pixel sj+1,2k+1 as the average of its two immediate level-j neighbors sj,k and sj,k+1, we calculate it as a weighted sum of its four immediate levelj neighbors sj,k-1, sj,k, sj,k+1, and sj,k+2. It is obvious that the two closer neighbors sj,k and sj,k+1 should be assigned more weight than the two extreme neighbors sj,k-1 and sj,k+2, but what should the weights be? The answer is given by Equation (7.45) (Section 7.25.4), which shows that the degree-3 (cubic) polynomial P(t) that interpolates four arbitrary points P1, P2, P3, and P4 is given by P(t) = (t3, t2, t, 1)... -4.5 13.5 -13.5 4.5 9.0 -22.5 18 -4.5 -5.5 9.0 -4.5 1.0 1.0 0 0 0 ... ... P1 P2 P3 P4 ... . (8.17) We are going to place the new odd-numbered pixel sj+1,2k+1 in the middle of the group of its four level-j neighbors, so it makes sense to assign it the value of the interpolating polynomial in the middle of its interval, i.e., P(0.5). Calculated from Equation (8.17), this value is P(0.5) = -0.0625P1 + 0.5625P2 + 0.5625P3 - 0.0625P4. 806 8. Wavelet Methods (Notice that the four weights add up to one. This is an example of barycentric functions. Also, see Exercise 7.44 for an explanation of the negative weights.) The subdivision rule in this case is [by analogy with Equation (8.16)] sj+1,2k = sj,k, sj+1,2k+1 = P3,j,k-1(0.5), (8.18) where the notation P3,j,k-1(t) indicates the degree-3 interpolating polynomial for the group of four level-j pixels that starts at sj,k-1. We define the order of this subdivision to be 4 (the number of interpolated pixels). Since linear subdivision interpolates two pixels, its order is 2. Figure 8.51 shows three levels of pixels generated by linear (8.51a) and by cubic (8.51b) subdivisions. (a) (b) j+2 j+1 j Figure 8.51: Linear and Cubic Subdivisions. It is now obvious how interpolating subdivision can be extended to higher orders. Select an even integer n, derive the degree-(n - 1) polynomial Pn-1,j,k-(n/2-1)(t) that will interpolate the n level-j pixels sj,k-(n/2-1), sj,k-n/2, sj,k-n/2+1, . . . , sj,k, sj,k+1, . . . , sj,k+n/2, 8.11 The Lifting Scheme 807 calculate the midpoint Pn-1,j,k-(n/2-1)(0.5), and generate the level-(j + 1) pixels according to sj+1,2k = sj,k, sj+1,2k+1 = Pn-1,j,k-(n/2-1)(0.5), (8.19)  Exercise 8.14: Compute the midpoint P(0.5) of the degree-5 interpolating polynomial for six points P1 through P6 as a function of the points. 8.11.3 Scaling Functions The scaling functions .j,k(x) have been mentioned in Section 8.6, in connection with the Haar transform. Here we show how they are constructed for an interpolating subdivision. Each coefficient sj,k computed in level j has a scaling function .j,k(x) associated with it, which is defined as follows: Select values n, j, and k. Set pixel sj,k to 1 and all other sj,i to 0 (this can be expressed as sj,k = .0,k). Use a subdivision scheme (based on n points) to compute levels (j +1), (j +2), etc. Each level has twice the number of pixels as its predecessor. In the limit, we end up with an infinite number of pixels. We can view these pixels as real numbers and define the range of the scaling function .j,k(x) as these numbers. Each pair of values j and k defines a different scaling function .j,k(x), but we can intuitively see that the scaling functions depend on j and k in simple ways. The shape of .j,k(x) does not depend on k, since we calculate . by setting sj,k = 1 and all other sj,i = 0. Thus, the function .j,8(x) is a shifted copy of .j,7(x), a twice shifted copy of .j,6(x), etc. In general, we may write .j,k(x) = .j,0(x - k). To understand the dependence of .j,k(x) on j, the reader should recall the following sentence (from page 804): “The original elements s0,k of s0 are now stored in a in locations k·2j.” This implies that if we select a small value for j, we end up with a wide scaling function .j,k(x). In general, we have .j,k(x) = .0,0(2jx - k) def = .(2jx - k), implying that all the scaling functions for different values of j and k are translations and dilations (scaling) of .(x), the fundamental solution of the subdivision process. .(x) is shown in Figure 8.52 for n=2, 4, 6, and 8. The main properties of .(x) are compact support and smoothness (it can be nonzero only inside the interval [-(n - 1), n - 1]), but it also satisfies a refinement relation of the form .(x) = n l=-n hl.(2x - l), where hl are called the filter coefficients. A change of variables allows us to write the same refinement relation in the form .j,k(x) = l hl-2k.j+1,l(x). The odd-numbered filter coefficients are the coefficients of the interpolating polynomial at its midpoint Pn-1,j,k-(n/2-1)(0.5). The even-numbered coefficients are zero except 808 8. Wavelet Methods -4 -3 -2 -1 0 0.0 0.5 1.0 1 2 3 4 -0.5 -3 -2 -1 0 1 2 3 4 -4 -3 -2 -1 0 0.0 0.5 1.0 1 2 3 4 -0.5 -3 -2 -1 0 1 2 3 4 Figure 8.52: Scaling Functions .j,k(x) for n=2, 4, 6, and 8. for l = 0 (this property can be expressed as h2l = .0,l). The general expression for hk is hk = "k even, .k,0, k odd, (-1)d+k .2d-1 i=0 (i-d+1/2) (k+1/2)(d+k)!(d-k-1)! , where d = n/2. For linear subdivision, the filter coefficients are (1/2, 1, 1/2). For cubic interpolation the eight filter coefficients are (-1/16, 0, 9/16, 1, 9/16, 0,-1/16). The filter coefficients are useful, since they allow us to write the interpolating subdivision in the form sj+1,l = k hl-2ksj,k. 8.12 The IWT 809 8.12 The IWT The DWT is simple but has an important drawback, namely, it uses noninteger filter coefficients, which is why it produces noninteger transform coefficients. There are several ways to modify the basic DWT such that it produces integer transform coefficients. This section describes a simple integer wavelet transform (IWT) that can be used to decompose an image in any of the ways described in Section 8.10. The transform is reversible, i.e., the image can be fully reconstructed from the (integer) transform coefficients. This IWT can be used to compress the image either lossily (by quantizing the transform coefficients) or losslessly (by entropy coding the transform coefficients). The simple principle of this transform is illustrated here for the one-dimensional case. Given a data vector of N integers xi, where i = 0, 1, . . . , N -1, we define k = N/2 and we compute the transform vector yi by calculating the odd and even components of y separately. We first discuss the case where N is even. The N/2 odd components y2i+1 (where i = 0, 1, . . . , k - 1) are calculated as differences of the xi’s. They become the detail (high frequency) transform coefficients. Each of the even components y2i (where i varies in the same range [0, k - 1]) is calculated as a weighted average of five data items xi. These N/2 numbers become the low-frequency transform coefficients, and are normally transformed again, into N/4 low-frequency and N/4 high-frequency coefficients. The basic rule for the odd transform coefficients is y2i+1 = - 1 2x2i + x2i+1 - 1 2x2i+2, except the last coefficient, where i = k - 1, which is computed as the simple difference y2k-1 = x2k-1 - x2k-2. This can be summarized as y2i+1 = x2i+1 - (x2i + x2i+2)/2, for i = 0, 1, . . . , k - 2, x2i+1 - x2i, for i = k - 1. (8.20) The even transform coefficients are calculated as the weighted average y2i = - 1 8x2i-2 + 1 4x2i-1 + 3 4x2i + 1 4x2i+1 - 1 8x2i+2, except the first coefficient, where i = 0, which is calculated as y0 = 3 4x0 + 1 2x1 - 1 4x2. In practice, this calculation is done by computing each even coefficient y2i in terms of x2i and the two odd coefficients y2i-1 and y2i+1. This can be summarized by y2i = x2i + y2i+1/2, for i = 0, x2i + (y2i-1 + y2i+1)/4, for i = 1, 2, . . . , k - 1. (8.21) 810 8. Wavelet Methods The inverse transform is easy to figure out. It uses the transform coefficients yi to calculate data items zi that are identical to the original xi. It first computes the even elements z2i = y2i - y2i+1/2, for i = 0, y2i - (y2i-1 + y2i+1)/4, for i = 1, 2, . . . , k - 1, (8.22) and then the odd elements z2i+1 = y2i+1 + (z2i + z2i+2)/2, for i = 0, 1, . . . , k - 2, y2i+1 + z2i, for i = k - 1. (8.23) Now comes the interesting part. The transform coefficients calculated by Equations (8.20) and (8.21) are generally nonintegers, because of the divisions by 2 and 4. The same is true for the reconstructed data items of Equations (8.22) and (8.23). The main feature of the particular IWT described here is the use of truncation. Truncation, denoted by the “floor” symbols  and , is used to produce integer transform coefficients yi and also integer reconstructed data items zi. Equations (8.20) through (8.23) are modified to y2i+1 = x2i+1 - (x2i + x2i+2)/2, for i = 0, 1, . . . , k - 2, x2i+1 - x2i, for i = k - 1. y2i = x2i + y2i+1/2, for i = 0, x2i + (y2i-1 + y2i+1)/4, for i = 1, 2, . . . , k - 1. z2i = y2i - y2i+1/2, for i = 0, y2i - (y2i-1 + y2i+1)/4, for i = 1, 2, . . . , k - 1, (8.24) z2i+1 = y2i+1 + (z2i + z2i+2)/2, for i = 0, 1, . . . , k - 2, y2i+1 + z2i, for i = k - 1. Because of truncation, some information is lost when the yi are calculated. However, truncation is also used in the calculation of the zi, which restores the lost information. Thus, Equation (8.24) is a true forward and inverse IWT that reconstructs the original data items exactly.  Exercise 8.15: Given the data vector x = (112, 97, 85, 99, 114, 120, 77, 80), use Equation (8.24) to calculate its forward and inverse integer wavelet transforms. The same concepts can be applied to the case where the number N of data items is odd. We first define k by N = 2k + 1, then define the forward and inverse integer transforms by y2i+1 = x2i+1 - (x2i + x2i+2)/2, for i = 0, 1, . . . , k - 1, y2i =... x2i + y2i+1/2, for i = 0, x2i + (y2i-1 + y2i+1)4, for i = 1, 2, . . . , k - 1, x2i + y2i-1/2, for i = k. z2i =... y2i - y2i+1/2, for i = 0, y2i - (y2i-1 + y2i+1)/4, for i = 1, 2, . . . , k - 1, y2i - y2i-1/2, for i = k, z2i+1 = y2i+1 + (z2i + z2i+2)/2, for i = 0, 1, . . . , k - 1. 8.13 The Laplacian Pyramid 811 Notice that the IWT produces a vector yi where the detail coefficients and the weighted averages are interleaved. The algorithm should be modified to place the averages in the first half of y and the details in the second half. The extension of this transform to the two-dimensional case is obvious. The IWT is applied to the rows and the columns of the image using any of the image decomposition methods discussed in Section 8.10. 8.13 The Laplacian Pyramid The main feature of the Laplacian pyramid method [Burt and Adelson 83] is progressive compression. The decoder inputs the compressed stream section by section, and each section improves the appearance on the screen of the image-so-far. The method uses both prediction and transform techniques, but its computations are simple and local (i.e., there is no need to examine or use values that are far away from the current pixel). The name “Laplacian” comes from the field of image enhancement, where it is used to indicate operations similar to the ones used here. We start with a general description of the method. We denote by g0(i, j) the original image. A new, reduced image g1 is computed from g0 such that each pixel of g1 is a weighted sum of a group of 5Χ5 pixels of g0. Image g1 is computed [see Equation (8.25)] such that it has half the number of rows and half the number of columns of g0, so it is one-quarter the size of g0. It is a blurred (or lowpass filtered) version of g0. The next step is to expand g1 to an image g1,1 the size of g0 by interpolating pixel values [Equation (8.26)]. A difference image (also called an error image) L0 is calculated as the difference g0 - g1,1, and it becomes the bottom level of the Laplacian pyramid. The original image g0 can be reconstructed from L0 and g1,1, and also from L0 and g1. Since g1 is smaller than g1,1, it makes sense to write L0 and g1 on the compressed stream. The size of L0 equals that of g0, and the size of g1 is 1/4 of that, so it seems that we have generated expansion, but in fact, compression is achieved, because the error values in L0 are decorrelated to a high degree, and so are small (and therefore have small variance and low entropy) and can be represented with fewer bits than the original pixels in g0. In order to achieve progressive representation of the image, only L0 is written on the output, and the process is repeated on g1. A new, reduced image g2 is computed from g1 (the size of g2 is 1/16 that of g0). It is expanded to an image g2,1, and a new difference image L1 is computed as the difference g1 - g2,1 and becomes the next level, above L0, of the Laplacian pyramid. The final result is a sequence L0, L1,. . . , Lk-1, Lk where the first k items are difference images and the last one, Lk, is simply the (very small) reduced image gk. These items constitute the Laplacian pyramid. They are written on the compressed stream in reverse order, so the decoder inputs Lk first, uses it to display a small, blurred image, inputs Lk-1, reconstructs and displays gk-1 (which is four times bigger), and repeats until g0 is reconstructed and displayed. This process can be summarized by gk = Lk, gi = Li + Expand(gi+1) = Li + gi+1,1, for i = k - 1, k - 2, . . . , 2, 1, 0. 812 8. Wavelet Methods The user sees small, blurred images that grow and become sharper. The decoder can be modified to expand each intermediate image gi (before it is displayed) several times, by interpolating pixel values, until it gets to the size of the original image g0. This way, the user sees an image that progresses from very blurred to sharp, while remaining the same size. For example, image g3, which is 1/64th the size of g0, can be brought to that size by expanding it three times, yielding the chain g3,3 = Expand(g3,2) = Expand(g3,1) = Expand(g3). In order for all the intermediate gi and Li images to have well-defined dimensions, the original image g0 should have R = MR2M + 1 rows and C = MC2M + 1 columns, where MR, MC, and M are integers. Selecting, for example, MR = MC results in a square image. Image g1 has dimensions (MR2M-1 +1) Χ (MC2M-1 + 1), and image gp has dimensions (MR2M-p + 1) Χ (MC2M-p + 1). An example is MR = MC = 1 and M = 8. The dimensions of the original image are (28+1)Χ(28+1) = 257Χ257 and the reduced images g1 through g5 have dimensions (27 +1)Χ(27 +1) = 129Χ129, 65Χ65, 33 Χ 33, 17 Χ 17, and 9 Χ 9, respectively.  Exercise 8.16: Calculate the dimensions of the first six images g0 through g5 for the case MC = 3, MR = 4, and M = 5. We now turn to the details of reducing and expanding images. Reducing an image gp-1 to an image gp of dimensions Rp Χ Cp (where Rp = MR2M-p + 1, and Cp = MC2M-p + 1) is done by gp(i, j) = 2 m=-2 2 n=-2 w(m, n)gp-1(2i + m, 2j + n), (8.25) where i = 0, 1, . . . , Cp - 1, j = 0, 1, . . . , Rp - 1, and p (the level) varies from 1 to k - 1. Each pixel of gp is a weighted sum of 5Χ5 pixels of gp-1 with weights w(m, n) that are the same for all levels p. Figure 8.53a illustrates this process in one dimension. It shows how each pixel in a higher level of the pyramid is generated as a weighted sum of five pixels from the level below it, and how each level has (about) half the number of pixels of its predecessor. Figure 8.53b is in two dimensions. It shows how a pixel in a high level is obtained from 25 pixels located one level below. Some of the weights are also shown. Notice that in this case each level has about 1/4 the number of pixels of its predecessor. The weights w(m, n) (also called the generating kernel) are determined by first separating each in the form w(m, n) = ˆ w(m) ˆ w(n), where the functions ˆ w(m) should be normalized, i.e., 2 m=-2 ˆ w(m) = 1, and symmetric, ˆ w(m) = ˆw(-m). These two constraints are not enough to determine ˆ w(m), so we add a third one, called equal contribution, that demands that all the pixels at a given level contribute the same total weight (= 1/4) to pixels at the next higher 8.13 The Laplacian Pyramid 813 a a b b b b c c c c g0 g1 g2 w(-2,2) w(-1,0) w(0,0) w(1,-2) w(2,-1) w(2,2) (a) (b) Figure 8.53: Illustrating Reduction. level. If we set ˆ w(0) = a, ˆw(-1) = ˆ w(1) = b, and ˆ w(-2) = ˆ w(2) = c, then the three constraints are satisfied if w(0) = a, ˆ w(-1) = ˆ w(1) = 0.25, ˆ w(-2) = ˆ w(2) = 0.25 - a/2. (The reader should compare this to the discussion of interpolating polynomials in Section 7.25.4.) Experience recommends setting a = 0.6, which yields (see values of w(m, n) in Table 8.54) ˆ w(0) = 0.6, ˆ w(-2) = ˆ w(2) = 0.25, ˆ w(-1) = ˆ w(1) = 0.25 - 0.3 = -0.05. The same weights used in reducing images are also used in expanding them. Expanding an image gp of dimensions Rp ΧCp (where Rp = MR2M-p + 1, and Cp = MC2M-p +1) to an image gp,1 that has the same dimensions as gp-1 (i.e., is four times 814 8. Wavelet Methods bigger than gp) is done by gp,1(i, j) = 4 2 m=-2 2 n=-2 w(m, n)gp[i -m]/2, [j - n]/2, (8.26) where i = 0, 1, . . . , Cp - 1, j = 0, 1, . . . , Rp - 1, and the sums include only terms for which both (i -m)/2 and (j - n)/2 are integers. As an example we compute the single pixel g1,1(4, 5) = 4 2 m=-2 2 n=-2 w(m, n)g1[4 -m]/2, [5 - n]/2. Of the 25 terms of this sum, only six, namely those with m = -2, 0, 2 and n = -1, 1, satisfy the condition above and are included. The six terms correspond to (m, n) = (-2,-1), (-2, 1), (0,-1), (0, 1), (2,-1), (2, 1), and the sum is 4w(-2,-1)g1(3, 3) + w(-2, 1)g1(3, 2) + w(0,-1)g1(2, 3) + w(0, 1)g1(2, 2) + w(2,-1)g1(1, 3) + w(2, 1)g1(1, 2). -0.05 0.25 0.6 0.25 -0.05 -0.05 0.0025 -0.0125 -0.03 -0.01 0.0025 0.25 0.0625 0.15 0.0625 -0.0125 0.6 0.36 0.15 -0.03 0.25 0.0625 -0.0125 -0.05 0.0025 Table 8.54: Values of w(m, n) for a = 0.6. A lossy version of the Laplacian pyramid can be obtained by quantizing the values of each Li image before it is encoded and written on the compressed stream. 8.14 SPIHT 815 8.14 SPIHT Section 8.6 shows how the Haar transform can be applied several times to an image, creating regions (or subbands) of averages and details. The Haar transform is simple, and better compression can be achieved by other wavelet filters. It seems that different wavelet filters produce different results depending on the image type, but it is currently not clear what filter is the best for any given image type. Regardless of the particular filter used, the image is decomposed into subbands, such that lower subbands correspond to higher image frequencies (they are the highpass levels) and higher subbands correspond to lower image frequencies (lowpass levels), where most of the image energy is concentrated (Figure 8.55). This is why we can expect the detail coefficients to get smaller as we move from high to low levels. Also, there are spatial similarities among the subbands (Figure 8.17b). An image part, such as an edge, occupies the same spatial position in each subband. These features of the wavelet decomposition are exploited by the SPIHT (set partitioning in hierarchical trees) method [Said and Pearlman 96]. Frequencies Energy Coarse resolution Fine resolution LL4 LH1 LH2 LH3 HL1 HL2 HL3 HH1 HH2 HH3 Figure 8.55: Subbands and Levels in Wavelet Decomposition. SPIHT was designed for optimal progressive transmission, as well as for compression. One of the important features of SPIHT (perhaps a unique feature) is that at any point during the decoding of an image, the quality of the displayed image is the best that can be achieved for the number of bits input by the decoder up to that moment. Another important SPIHT feature is its use of embedded coding. This feature is defined as follows: If an (embedded coding) encoder produces two files, a large one of 816 8. Wavelet Methods size M and a small one of size m, then the smaller file is identical to the first m bits of the larger file. The following example aptly illustrates the meaning of this definition. Suppose that three users wait for you to send them a certain compressed image, but they need different image qualities. The first one needs the quality contained in a 10 Kb file. The image qualities required by the second and third users are contained in files of sizes 20 Kb and 50 Kb, respectively. Most lossy image compression methods would have to compress the same image three times, at different qualities, to generate three files with the right sizes. SPIHT, on the other hand, produces one file, and then three chunks—of lengths 10 Kb, 20 Kb, and 50 Kb, all starting at the beginning of that file—can be sent to the three users, thereby satisfying their needs. We start with a general description of SPIHT. We denote the pixels of the original image p by pi,j . Any set T of wavelet filters can be used to transform the pixels to wavelet coefficients (or transform coefficients) ci,j . These coefficients constitute the transformed image c. The transformation is denoted by c = T(p). In a progressive transmission method, the decoder starts by setting the reconstruction image ˆc to zero. It then inputs (encoded) transform coefficients, decodes them, and uses them to generate an improved reconstruction image ˆc, which, in turn, is used to produce a better image ˆp. We can summarize this operation by ˆp = T-1(ˆc). The main aim in progressive transmission is to transmit the most important image information first. This is the information that results in the largest reduction of the distortion (the difference between the original and the reconstructed images). SPIHT uses the mean squared error (MSE) distortion measure [Equation (7.2)] Dmse(p - ˆp) = |p - ˆp|2 N = 1 N i j (pi,j - ˆpi,j)2, where N is the total number of pixels. An important consideration in the design of SPIHT is the fact that this measure is invariant to the wavelet transform, a feature that allows us to write Dmse(p - ˆp) = Dmse(c - ˆc) = |p - ˆp|2 N = 1 N i j (ci,j - ˆci,j)2. (8.27) Equation (8.27) shows that the MSE decreases by |ci,j |2/N when the decoder receives the transform coefficient ci,j (we assume that the decoder receives the exact value of the coefficient, i.e., there is no loss of precision due to limitations imposed by computer arithmetic). It is now clear that the largest coefficients ci,j (largest in absolute value, regardless of their signs) contain the information that reduces the MSE distortion most, so a progressive encoder should send those coefficients first. This is an important principle of SPIHT. Another principle is based on the observation that the most significant bits of a binary integer whose value is close to maximum tend to be ones. This suggests that the most significant bits contain the most important image information, and that they should be sent to the decoder first (or written first on the compressed stream). 8.14 SPIHT 817 The progressive transmission method used by SPIHT incorporates these two principles. SPIHT sorts the coefficients and transmits their most significant bits first. To simplify the description, we first assume that the sorting information is explicitly transmitted to the decoder; the next section shows an efficient way to code this information. We now show how the SPIHT encoder uses these principles to progressively transmit the wavelet coefficients to the decoder (or write them on the compressed stream), starting with the most important information. We assume that a wavelet transform has already been applied to the image (SPIHT is a coding method, so it can work with any wavelet transform) and that the transformed coefficients ci,j are already stored in memory. The coefficients are sorted (ignoring their signs), and the sorting information is contained in an array m such that array element m(k) contains the (i, j) coordinates of a coefficient ci,j , and such that |cm(k)| . |cm(k+1)| for all values of k. Table 8.56 lists hypothetical values of 16 coefficients. Each is shown as a 16-bit number where the most significant bit (bit 15) is the sign and the remaining 15 bits (numbered 14 through 0, top to bottom) constitute the magnitude. The first coefficient cm(1) = c2,3 is s1aci . . . r (where s, a, etc., are bits). The second one cm(2) = c3,4 is s1bdj . . . s, and so on. k 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 sign s s s s s s s s s s s s s s s s msb 14 1 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 13 a b 1 1 1 1 0 0 0 0 0 0 0 0 0 0 12 c d e f g h 1 1 1 0 0 0 0 0 0 0 11 i j k l m n o p q 1 0 0 0 0 0 0 ... ... ... ... lsb 0 r s t u v w x y . . . . . . . . . z m(k) = i, j 2,3 3,4 3,2 4,4 1,2 3,1 3,3 4,2 4, 1 . . . . . . . . . 4, 3 Table 8.56: Transform Coefficients Ordered by Absolute Magnitudes. The sorting information that the encoder has to transmit is the sequence m(k), or (2, 3), (3, 4), (3, 2), (4, 4), (1, 2), (3, 1), (3, 3), (4, 2), . . . , (4, 3). In addition, it has to transmit the 16 signs, and the 16 coefficients in order of significant bits. A direct transmission would send the 16 numbers ssssssssssssssss, 1100000000000000, ab11110000000000, cdefgh1110000000, ijklmnopq1000000, . . . , rstuvwxy . . . z, but this is clearly wasteful. Instead, the encoder goes into a loop, where in each iteration it performs a sorting step and a refinement step. In the first iteration it transmits the number l = 2 (the number of coefficients ci,j in our example that satisfy 214 . |ci,j | < 215) followed by the two pairs of coordinates (2, 3) and (3, 4) and by the signs 818 8. Wavelet Methods of the first two coefficients. This is done in the first sorting pass. This information enables the decoder to construct approximate versions of the 16 coefficients as follows: Coefficients c2,3 and c3,4 are constructed as the 16-bit numbers s100 . . . 0. The remaining 14 coefficients are constructed as all zeros. This is how the most significant bits of the largest coefficients are transmitted to the decoder first. The next step of the encoder is the refinement pass, but this is not performed in the first iteration. In the second iteration the encoder performs both passes. In the sorting pass it transmits the number l = 4 (the number of coefficients ci,j in our example that satisfy 213 . |ci,j | < 214), followed by the four pairs of coordinates (3, 2), (4, 4), (1, 2), and (3, 1) and by the signs of the four coefficients. In the refinement step it transmits the two bits a and b. These are the 14th most significant bits of the two coefficients transmitted in the previous iteration. The information received so far enables the decoder to improve the 16 approximate coefficients constructed in the previous iteration. The first six become c2,3 = s1a0 . . . 0, c3,4 = s1b0 . . . 0, c3,2 = s0100 . . . 0, c4,4 = s0100 . . . 0, c1,2 = s0100 . . . 0, c3,1 = s0100 . . . 0, and the remaining 10 coefficients are not changed.  Exercise 8.17: Perform the sorting and refinement passes of the next (third) iteration. The main steps of the SPIHT encoder should now be easy to understand. They are as follows: Step 1: Given an image to be compressed, perform its wavelet transform using any suitable wavelet filter, decompose it into transform coefficients ci,j , and represent the resulting coefficients with a fixed number of bits. (In the discussion that follows we use the terms pixel and coefficient interchangeably.) We assume that the coefficients are represented as 16-bit signed-magnitude numbers. The leftmost bit is the sign, and the remaining 15 bits are the magnitude. (Notice that the sign-magnitude representation is different from the 2’s complement method, which is used by computer hardware to represent signed numbers.) Such numbers can have values from -(215 - 1) to 215 - 1. Set n to log2 maxi,j(ci,j). In our case n will be set to log2(215 - 1) = 14. Step 2: Sorting pass: Transmit the number l of coefficients ci,j that satisfy 2n . |ci,j | < 2n+1. Follow with the l pairs of coordinates and the l sign bits of those coefficients. Step 3: Refinement pass: Transmit the nth most significant bit of all the coefficients satisfying |ci,j| . 2n+1. These are the coefficients that were selected in previous sorting passes (not including the immediately preceding sorting pass). Step 4: Iterate: Decrement n by 1. If more iterations are needed (or desired), go back to Step 2. The last iteration is normally performed for n = 0, but the encoder can stop earlier, in which case the least important image information (some of the least significant bits of all the wavelet coefficients) will not be transmitted. This is the natural lossy option of SPIHT. It is equivalent to scalar quantization, but it produces better results than what is usually achieved with scalar quantization, since the coefficients are transmitted in sorted order. An alternative is for the encoder to transmit the entire image (i.e., 8.14 SPIHT 819 all the bits of all the wavelet coefficients) and the decoder can stop decoding when the reconstructed image reaches a certain quality. This quality can either be predetermined by the user or automatically determined by the decoder at run time. 8.14.1 Set Partitioning Sorting Algorithm The method as described so far is simple, since we have assumed that the coefficients had been sorted before the loop started. In practice, the image may have 1KΧ1K pixels or more; there may be more than a million coefficients, so sorting all of them is too slow. Instead of sorting the coefficients, SPIHT uses the fact that sorting is done by comparing two elements at a time, and each comparison results in a simple yes/no result. Therefore, if both encoder and decoder use the same sorting algorithm, the encoder can simply send the decoder the sequence of yes/no results, and the decoder can use those to duplicate the operations of the encoder. This is true not just for sorting but for any algorithm based on comparisons or on any type of branching. The actual algorithm used by SPIHT is based on the realization that there is really no need to sort all the coefficients. The main task of the sorting pass in each iteration is to select those coefficients that satisfy 2n . |ci,j | < 2n+1. This task is divided into two parts. For a given value of n, if a coefficient ci,j satisfies |ci,j| . 2n, then we say that it is significant; otherwise, it is called insignificant. In the first iteration, relatively few coefficients will be significant, but their number increases from iteration to iteration, because n keeps getting decremented. The sorting pass has to determine which of the significant coefficients satisfies |ci,j | < 2n+1 and transmit their coordinates to the decoder. This is an important part of the algorithm used by SPIHT. The encoder partitions all the coefficients into a number of sets Tk and performs the significance test max (i,j).Tk |ci,j| . 2n ? on each set Tk. The result may be either “no” (all the coefficients in Tk are insignificant, so Tk itself is considered insignificant) or “yes” (some coefficients in Tk are significant, so Tk itself is significant). This result is transmitted to the decoder. If the result is “yes,” then Tk is partitioned by both encoder and decoder, using the same rule, into subsets and the same significance test is performed on all the subsets. This partitioning is repeated until all the significant sets are reduced to size 1 (i.e., they contain one coefficient each, and that coefficient is significant). This is how the significant coefficients are identified by the sorting pass in each iteration. The significance test performed on a set T can be summarized by Sn(T) = 1, max(i,j).T |ci,j| . 2n, 0, otherwise. (8.28) The result, Sn(T), is a single bit that is transmitted to the decoder. The result of each significance test becomes a single bit written on the compressed stream, which is why the number of tests should be minimized. To achieve this goal, the sets should be created and partitioned such that sets expected to be significant will be large and sets that are expected to be insignificant will contain just one element. 820 8. Wavelet Methods 8.14.2 Spatial Orientation Trees The sets Tk are created and partitioned using a special data structure called a spatial orientation tree. This structure is defined in a way that exploits the spatial relationships between the wavelet coefficients in the different levels of the subband pyramid. Experience has shown that the subbands in each level of the pyramid exhibit spatial similarity (Figure 8.17b). Any special features, such as a straight edge or a uniform region, are visible in all the levels at the same location. The spatial orientation trees are illustrated in Figure 8.57a,b for a 16Χ16 image. The figure shows two levels, level 1 (the highpass) and level 2 (the lowpass). Each level is divided into four subbands. Subband LL2 (the lowpass subband) is divided into four groups of 2Χ2 coefficients each. Figure 8.57a shows the top-left group, and Figure 8.57b shows the bottom-right group. In each group, each of the four coefficients (except the top-left one, marked in gray) becomes the root of a spatial orientation tree. The arrows show examples of how the various levels of these trees are related. The thick arrows indicate how each group of 4Χ4 coefficients in level 2 is the parent of four such groups in level 1. In general, a coefficient at location (i, j) in the image is the parent of the four coefficients at locations (2i, 2j), (2i + 1, 2j), (2i, 2j + 1), and (2i + 1, 2j + 1). The roots of the spatial orientation trees of our example are located in subband LL2 (in general, they are located in the top-left LL subband, which can be of any size), but any wavelet coefficient, except the gray ones on level 1 (also except the leaves), can be considered the root of some spatial orientation subtree. The leaves of all those trees are located on level 1 of the subband pyramid. In our example, subband LL2 is of size 4Χ4, so it is divided into four 2Χ2 groups, and three of the four coefficients of a group become roots of trees. Thus, the number of trees in our example is 12. In general, the number of trees is 3/4 the size of the highest LL subband. Each of the 12 roots in subband LL2 in our example is the parent of four children located on the same level. However, the children of these children are located on level 1. This is true in general. The roots of the trees are located on the highest level, and their children are on the same level, but from then on, the four children of a coefficient on level k are themselves located on level k - 1. We use the terms offspring for the four children of a node, and descendants for the children, grandchildren, and all their descendants. The set partitioning sorting algorithm uses the following four sets of coordinates: 1. O(i, j): the set of coordinates of the four offspring of node (i, j). If node (i, j) is a leaf of a spatial orientation tree, then O(i, j) is empty. 2. D(i, j): the set of coordinates of the descendants of node (i, j). 3. H(i, j): the set of coordinates of the roots of all the spatial orientation trees (3/4 of the wavelet coefficients in the highest LL subband). 4. L(i, j): The difference set D(i, j)-O(i, j). This set contains all the descendants of tree node (i, j), except its four offspring. The spatial orientation trees are used to create and partition the sets Tk. The set partitioning rules are as follows: 1. The initial sets are {(i, j)} and D(i, j), for all (i, j) . H (i.e., for all roots of the spatial orientation trees). In our example there are 12 roots, so there will initially be 24 8.14 SPIHT 821 1 2 3 4 LH2 LH1 HL2 HH2 HL1 HH1 LL2 1 2 3 4 LH2 LL2 LH1 HL2 HL1 HH1 (a) (b) 1 2 3 3 4 2 3 Figure 8.57: Spatial Orientation Trees in SPIHT. sets: 12 sets, each containing the coordinates of one root, and 12 more sets, each with the coordinates of all the descendants of one root. 2. If set D(i, j) is significant, then it is partitioned into L(i, j) plus the four singleelement sets with the four offspring of (i, j). In other words, if any of the descendants of node (i, j) is significant, then its four offspring become four new sets and all its other descendants become another set (to be significance tested in rule 3). 3. If L(i, j) is significant, then it is partitioned into the four sets D(k, l), where (k, l) are the offspring of (i, j). Once the spatial orientation trees and the set partitioning rules are understood, the coding algorithm can be described. 8.14.3 SPIHT Coding It is important to have the encoder and decoder test sets for significance in the same way. The coding algorithm therefore uses three lists called list of significant pixels (LSP), list of insignificant pixels (LIP), and list of insignificant sets (LIS). These are lists of coordinates (i, j) that in the LIP and LSP represent individual coefficients, and in the LIS represent either the set D(i, j) (a type A entry) or the set L(i, j) (a type B entry). The LIP contains coordinates of coefficients that were insignificant in the previous sorting pass. In the current pass they are tested, and those that test significant are moved to the LSP. In a similar way, sets in the LIS are tested in sequential order, and when a set is found to be significant, it is removed from the LIS and is partitioned. The new subsets with more than one coefficient are placed back in the LIS, to be tested later, and the subsets with one element are tested and appended to the LIP or the LSP, depending on the results of the test. The refinement pass transmits the nth most significant bit of the entries in the LSP. Figure 8.58 shows this algorithm in detail. Figure 8.59 is a simplified version, for readers who are intimidated by too many details. The decoder executes the detailed algorithm of Figure 8.58. It always works in lockstep with the encoder, but the following notes shed more light on its operation: 822 8. Wavelet Methods 1. Initialization: Set n to log2 maxi,j(ci,j) and transmit n. Set the LSP to empty. Set the LIP to the coordinates of all the roots (i, j) . H. Set the LIS to the coordinates of all the roots (i, j) . H that have descendants. 2. Sorting pass: 2.1 for each entry (i, j) in the LIP do: 2.1.1 output Sn(i, j); 2.1.2 if Sn(i, j) = 1, move (i, j) to the LSP and output the sign of ci,j ; 2.2 for each entry (i, j) in the LIS do: 2.2.1 if the entry is of type A, then • output Sn(D(i, j)); • if Sn(D(i, j)) = 1, then * for each (k, l) . O(i, j) do: · output Sn(k, l); · if Sn(k, l) = 1, add (k, l) to the LSP, output the sign of ck,l; · if Sn(k, l) = 0, append (k, l) to the LIP; * if L(i, j) = 0, move (i, j) to the end of the LIS, as a type-B entry, and go to step 2.2.2; else, remove entry (i, j) from the LIS; 2.2.2 if the entry is of type B, then • output Sn(L(i, j)); • if Sn(L(i, j)) = 1, then * append each (k, l) . O(i, j) to the LIS as a type-A entry: * remove (i, j) from the LIS: 3. Refinement pass: for each entry (i, j) in the LSP, except those included in the last sorting pass (the one with the same n), output the nth most significant bit of |ci,j |; 4. Loop: decrement n by 1 and go to step 2 if needed. Figure 8.58: The SPIHT Coding Algorithm. 1. Set the threshold. Set LIP to all root nodes coefficients. Set LIS to all trees (assign type D to them). Set LSP to an empty set. 2. Sorting pass: Check the significance of all coefficients in LIP: 2.1 If significant, output 1, output a sign bit, and move the coefficient to the LSP. 2.2 If not significant, output 0. 3. Check the significance of all trees in the LIS according to the type of tree: 3.1 For a tree of type D: 3.1.1 If it is significant, output 1, and code its children: 3.1.1.1 If a child is significant, output 1, then a sign bit, add it to the LSP 3.1.1.2 If a child is insignificant, output 0 and add the child to the end of LIP. 3.1.1.3 If the children have descendants, move the tree to the end of LIS as type L, otherwise remove it from LIS. 3.1.2 If it is insignificant, output 0. 3.2 For a tree of type L: 3.2.1 If it is significant, output 1, add each of the children to the end of LIS as an entry of type D and remove the parent tree from the LIS. 3.2.2 If it is insignificant, output 0. 4. Loop: Decrement the threshold and go to step 2 if needed. Figure 8.59: A Simplified SPIHT Coding Algorithm. 8.14 SPIHT 823 1. Step 2.2 of the algorithm evaluates all the entries in the LIS. However, step 2.2.1 appends certain entries to the LIS (as type-B) and step 2.2.2 appends other entries to the LIS (as type-A). It is important to realize that all these entries are also evaluated by step 2.2 in the same iteration. 2. The value of n is decremented in each iteration, but there is no need to bring it all the way to zero. The loop can stop after any iteration, resulting in lossy compression. Normally, the user specifies the number of iterations, but it is also possible to have the user specify the acceptable amount of distortion (in units of MSE), and the encoder can use Equation (8.27) to decide when to stop the loop. 3. The encoder knows the values of the wavelet coefficients ci,j and uses them to calculate the bits Sn (Equation (8.28)), which it transmits (i.e., writes on the compressed stream). These bits are input by the decoder, which uses them to calculate the values of ci,j . The algorithm executed by the decoder is that of Figure 8.58 but with the word “output” changed to “input.” 4. The sorting information, previously denoted by m(k), is recovered when the coordinates of the significant coefficients are appended to the LSP in steps 2.1.2 and 2.2.1. This implies that the coefficients indicated by the coordinates in the LSP are sorted according to log2 |cm(k)| . log2 |cm(k+1)|, for all values of k. The decoder recovers the ordering because its three lists (LIS, LIP, and LSP) are updated in the same way as those of the encoder (remember that the decoder works in lockstep with the encoder). When the decoder inputs data, its three lists are identical to those of the encoder at the moment it (the encoder) output that data. 5. The encoder starts with the wavelet coefficients ci,j ; it never gets to “see” the actual image. The decoder, however, has to display the image and update the display in each iteration. In each iteration, when the coordinates (i, j) of a coefficient ci,j are moved to the LSP as an entry, it is known (to both encoder and decoder) that 2n . |ci,j | < 2n+1. As a result, the best value that the decoder can give the coefficient ˆci,j that is being reconstructed is midway between 2n and 2n+1 = 2Χ2n. Thus, the decoder sets ˆci,j = ±1.5Χ2n (the sign of ˆci,j is input by the decoder just after the insertion). During the refinement pass, when the decoder inputs the actual value of the nth bit of ci,j, it improves the value 1.5Χ2n by either adding 2n-1 to it (if the input bit was a 1) or subtracting 2n-1 from it (if the input bit was a 0). This way, the decoder can improve the appearance of the image (or, equivalently, reduce the distortion) during both the sorting and refinement passes. It is possible to improve the performance of SPIHT by entropy coding the encoder’s output, but experience shows that the added compression gained in this way is minimal and does not justify the additional expense of both encoding and decoding time. It turns out that the signs and the individual bits of coefficients output in each iteration are uniformly distributed, so entropy coding them does not produce any compression. The bits Sn(i, j) and Sn(D(i, j)), on the other hand, are distributed nonuniformly and may gain from such coding. 824 8. Wavelet Methods 8.14.4 Example We assume that a 4Χ4 image has already been transformed, and the 16 coefficients are stored in memory as 6-bit signed-magnitude numbers (one sign bit followed by five magnitude bits). They are shown in Figure 8.60, together with the single spatial orientation tree. The coding algorithm initializes LIP to the one-element set {(1, 1)}, the LIS to the set {D(1, 1)}, and the LSP to the empty set. The largest coefficient is 18, so n is set to log2 18 = 4. The first two iterations are shown. 18 6 8 -7 3 -5 13 1 2 1-6 3 2 -2 4-2 18 6 3 -5 8 (1,1) 13 (1,1) -7 (1,1) 1 (1,1) 2 (1,1) 1 (1,1) 2 (1,1) -2 (1,1) 4 (1,1) -6 (1,1) 3 (1,1) -2 (1,1) Figure 8.60: Sixteen Coefficients and One Spatial Orientation Tree. Sorting Pass 1: 2n = 16. Is (1,1) significant? yes: output a 1. LSP = {(1, 1)}, output the sign bit: 0. Is D(1, 1) significant? no: output a 0. LSP = {(1, 1)}, LIP = {}, LIS = {D(1, 1)}. Three bits output. Refinement pass 1: no bits are output (this pass deals with coefficients from sorting pass n - 1). Decrement n to 3. Sorting Pass 2: 2n = 8. Is D(1, 1) significant? yes: output a 1. Is (1, 2) significant? no: output a 0. Is (2, 1) significant? no: output a 0. Is (2, 2) significant? no: output a 0. LIP = {(1, 2), (2, 1), (2, 2)}, LIS = {L(1, 1)}. Is L(1, 1) significant? yes: output a 1. LIS = {D(1, 2),D(2, 1),D(2, 2)}. Is D(1, 2) significant? yes: output a 1. Is (1, 3) significant? yes: output a 1. LSP = {(1, 1), (1, 3)}, output sign bit: 1. Is (2, 3) significant? yes: output a 1. LSP = {(1, 1), (1, 3), (2, 3)}, output sign bit: 1. Is (1, 4) significant? no: output a 0. Is (2, 4) significant? no: output a 0. LIP = {(1, 2), (2, 1), (2, 2), (1, 4), (2, 4)}, 8.14 SPIHT 825 LIS = {D(2, 1),D(2, 2)}. Is D(2, 1) significant? no: output a 0. Is D(2, 2) significant? no: output a 0. LIP = {(1, 2), (2, 1), (2, 2), (1, 4), (2, 4)}, LIS = {D(2, 1),D(2, 2)}, LSP = {(1, 1), (1, 3), (2, 3)}. Fourteen bits output. Refinement pass 2: After iteration 1, the LSP included entry (1, 1), whose value is 18 = 100102. One bit is output. Sorting Pass 3: 2n = 4. Is (1, 2) significant? yes: output a 1. LSP = {(1, 1), (1, 3), (2, 3), (1, 2)}, output a sign bit: 1. Is (2, 1) significant? no: output a 0. Is (2, 2) significant? yes: output a 1. LSP = {(1, 1), (1, 3), (2, 3), (1, 2), (2, 2)}, output a sign bit: 0. Is (1, 4) significant? yes: output a 1. LSP = {(1, 1), (1, 3), (2, 3), (1, 2), (2, 2), (1, 4)}, output a sign bit: 1. Is (2, 4) significant? no: output a 0. LIP = {(2, 1), (2, 4)}. Is D(2, 1) significant? no: output a 0. Is D(2, 2) significant? yes: output a 1. Is (3, 3) significant? yes: output a 1. LSP = {(1, 1), (1, 3), (2, 3), (1, 2), (2, 2), (1, 4), (3, 3)}, output a sign bit: 0. Is (4, 3) significant? yes: output a 1. LSP = {(1, 1), (1, 3), (2, 3), (1, 2), (2, 2), (1, 4), (3, 3), (4, 3)}, output a sign bit: 1. Is (3, 4) significant? no: output a 0. LIP = {(2, 1), (2, 4), (3, 4)}. Is (4, 4) significant? no: output a 0. LIP = {(2, 1), (2, 4), (3, 4), (4, 4)}. LIP = {(2, 1), (3, 4), (3, 4), (4, 4)}, LIS = {D(2, 1)}, LSP = {(1, 1), (1, 3), (2, 3), (1, 2), (2, 2), (1, 4), (3, 3), (4, 3)}. Sixteen bits output. Refinement Pass 3: After iteration 2, the LSP included entries (1, 1), (1, 3), and (2, 3), whose values are 18 = 100102, 8 = 10002, and 13 = 11012. Three bits are output After two iterations, a total of 37 bits has been output. 8.14.5 QTCQ Closely related to SPIHT, the QTCQ (quadtree classification and trellis coding quantization) method [Banister and Fischer 99] uses fewer lists than SPIHT and explicitly forms classes of the wavelet coefficients for later quantization by means of the ACTCQ and TCQ (arithmetic and trellis coded quantization) algorithms of [Joshi, Crump, and Fischer 93]. 826 8. Wavelet Methods The method uses the spatial orientation trees originally developed by SPIHT. This type of tree is a special case of a quadtree (Section 7.34). The encoding algorithm is iterative. In the nth iteration, if any element of this quadtree is found to be significant, then the four highest elements in the tree are defined to be in class n. They also become roots of four new quadtrees. Each of the four new trees is tested for significance, moving down each tree until all the significant elements are found. All the wavelet coefficients declared to be in class n are stored in a list of pixels (LP). The LP is initialized with all the wavelet coefficients in the lowest frequency subband (LFS). The test for significance is performed by the function ST (k), which is defined by ST (k) = 1, max(i,j).k |Ci,j| . T, 0, otherwise, where T is the current threshold for significance and k is a tree of wavelet coefficients. The QTCQ encoding algorithm uses this test, and is listed in Figure 8.61. The QTCQ decoder is similar. All the outputs in Figure 8.61 should be replaced by inputs, and ACTCQ encoding should be replaced by ACTCQ decoding. 1. Initialization: Initialize LP with all Ci,j in LFS, Initialize LIS with all parent nodes, Output n = log2(max |Ci,j |/q). Set the threshold T = q2n, where q is a quality factor. 2. Sorting: for each node k in LIS do output ST (k) if ST (k) = 1 then for each child of k do move coefficients to LP add to LIS as a new node endfor remove k from LIS endif endfor 3. Quantization: For each element in LP, quantize and encode using ACTCQ. (use TCQ step size . = . · q). 4. Update: Remove all elements in LP. Set T = T/2. Go to step 2. Figure 8.61: QTCQ Encoding. The QTCQ implementation, as described in [Banister and Fischer 99], does not transmit the image progressively, but the authors claim that this property can be added to it. 8.15 CREW 827 8.15 CREW The CREW method (compression with reversible embedded wavelets) was developed in 1994 by A. Zandi at Ricoh Silicon Valley for the high-quality lossy and lossless compression of medical images. It was later realized that he independently developed a method very similar to SPIHT (Section 8.14), which is why the details of CREW are not described here. The interested reader is referred to [Zandi et al. 95], but more recent and detailed descriptions can be found at [CREW 00]. 8.16 EZW The SPIHT method is in some ways an extension of EZW, so this method, whose full name is “embedded coding using zerotrees of wavelet coefficients,” is described here by outlining its principles and showing an example. Some of the details, such as the relation between parents and descendants in a spatial orientation tree, and the meaning of the term “embedded,” are described in Section 8.14. The EZW method, as implemented in practice, starts by performing the 9-tap symmetric quadrature mirror filter (QMF) wavelet transform [Adelson et al. 87]. The main loop is then repeated for values of the threshold that are halved at the end of each iteration. The threshold is used to calculate a significance map of significant and insignificant wavelet coefficients. Zerotrees are used to represent the significance map in an efficient way. The main steps are as follows: 1. Initialization: Set the threshold T to the smallest power of 2 that is greater than max(i,j) |ci,j |/2, where ci,j are the wavelet coefficients. 2. Significance map coding: Scan all the coefficients in a predefined way and output a symbol when |ci,j | > T. When the decoder inputs this symbol, it sets ci,j = ±1.5T. 3. Refinement: Refine each significant coefficient by sending one more bit of its binary representation. When the decoder receives this, it increments the current coefficient value by ±0.25T. 4. Set T = T/2, and go to step 2 if more iterations are needed. A wavelet coefficient ci,j is considered insignificant with respect to the current threshold T if |ci,j| . T. The zerotree data structure is based on the following wellknown experimental result: If a wavelet coefficient at a coarse scale (i.e., high in the image pyramid) is insignificant with respect to a given threshold T, then all of the coefficients of the same orientation in the same spatial location at finer scales (i.e., located lower in the pyramid) are very likely to be insignificant with respect to T. In each iteration, all the coefficients are scanned in the order shown in Figure 8.62a. This guarantees that when a node is visited, all its parents will already have been scanned. The scan starts at the lowest frequency subband LLn, continues with subbands HLn, LHn, and HHn, and drops to level n - 1, where it scans HLn-1, LHn-1, and HHn-1. Each subband is fully scanned before the algorithm proceeds to the next subband. Each coefficient visited in the scan is classified as a zerotree root (ZTR), an isolated zero (IZ), positive significant (POS), or negative significant (NEG). A zerotree root is 828 8. Wavelet Methods a coefficient that is insignificant and all its descendants (in the same spatial orientation tree) are also insignificant. Such a coefficient becomes the root of a zerotree. It is encoded with a special symbol (denoted by ZTR), and the important point is that its descendants don’t have to be encoded in the current iteration. When the decoder inputs a ZTR symbol, it assigns a zero value to the coefficients and to all its descendants in the spatial orientation tree. Their values get improved (refined) in subsequent iterations. An isolated zero is a coefficient that is insignificant but has some significant descendants. Such a coefficient is encoded with the special IZ symbol. The other two classes are coefficients that are significant and are positive or negative. The flowchart of Figure 8.62b illustrates this classification. Notice that a coefficient is classified into one of five classes, but the fifth class (a zerotree node) is not encoded. (a) HH1 HL1 LH1 LH2 HH2 HL2 (b) plus minus yes yes yes no no no do not code insignificant, IZ ZTR POS NEG HH3 HL3 LH3 LL3 coefficient does descend from a zerotree root ? is coefficient significant ? coefficient does descendants have significant ? sign ? Figure 8.62: (a) Scanning a Zerotree. (b) Classifying a Coefficient. Coefficients in the lowest pyramid level don’t have any children, so they cannot be the roots of zerotrees. Thus, they are classified into isolated zero, positive significant, or negative significant. The zerotree can be viewed as a structure that helps to find insignificance. Most methods that try to find structure in an image try to find significance. The IFS method of Section 7.39, for example, tries to locate image parts that are similar up to size and/or transformation, and this is much harder than to locate parts that are insignificant relative to other parts. Two lists are used by the encoder (and also by the decoder, which works in lockstep) in the scanning process. The dominant list contains the coordinates of the coefficients that have not been found to be significant. They are stored in the order scan, by pyramid levels, and within each level by subbands. The subordinate list contains the magnitudes 8.16 EZW 829 (not coordinates) of the coefficients that have been found to be significant. Each list is scanned once per iteration. An iteration consists of a dominant pass followed by a subordinate pass. In the dominant pass, coefficients from the dominant list are tested for significance. If a coefficient is found significant, then (1) its sign is determined, (2) it is classified as either POS or NEG, (3) its magnitude is appended to the subordinate list, and (4) it is set to zero in memory (in the array containing all the wavelet coefficients). The last step is done so that the coefficient does not prevent the occurrence of a zerotree in subsequent dominant passes at smaller thresholds. Imagine that the initial threshold is T = 32. When a coefficient ci,j = 63 is encountered in the first iteration, it is found to be significant. Since it is positive, it is encoded as POS. When the decoder inputs this symbol, it does not know its value, but it knows that the coefficient is positive significant, i.e., it satisfies ci,j > 32. The decoder also knows that ci,j . 64 = 2Χ32, so the best value that the decoder can assign the coefficient is (32 + 64)/2 = 48. The coefficient is then set to 0, so subsequent iterations will not identify it as significant. We can think of the threshold T as an indicator that specifies a bit position. In each iteration, the threshold indicates the next less significant bit position. We can also view T as the current quantization width. In each iteration that width is divided by 2, so another less significant bit of the coefficients becomes known. During a subordinate pass, the subordinate list is scanned and the encoder outputs a 0 or a 1 for each coefficient to indicate to the decoder how the magnitude of the coefficient should be improved. In the example of ci,j = 63, the encoder transmits a 1, indicating to the decoder that the actual value of the coefficient is greater than 48. The decoder uses that information to improve the magnitude of the coefficient from 48 to (48+64)/2 = 56. If a 0 had been transmitted, the decoder would have refined the value of the coefficient to (32 + 48)/2 = 40. The string of bits generated by the encoder during the subordinate pass is entropy encoded using a custom version of adaptive arithmetic coding (Section 5.10). At the end of the subordinate pass, the encoder (and, in lockstep, also the decoder) sorts the magnitudes in the subordinate list in decreasing order. The encoder stops the loop when a certain condition is met. The user may, for example, specify the desired bitrate (number of bits per pixel). The encoder knows the image size (number of pixels), so it knows when the desired bitrate has been reached or exceeded. The advantage is that the compression ratio is known (in fact, it is determined by the user) in advance. The downside is that too much information may be lost if the compression ratio is too high, thereby leading to a poorly reconstructed image. It is also possible for the encoder to stop in the midst of an iteration, when the exact bitrate specified by the user has been reached. However, in such a case the last codeword may be incomplete, leading to wrong decoding of one coefficient. The user may also specify the bit budget (the size of the compressed stream) as a stopping condition. This is similar to specifying the bitrate. An alternative is for the user to specify the maximum acceptable distortion (the difference between the original and the compressed image). In such a case, the encoder iterates until the threshold becomes 1, and the decoder stops decoding when the maximum acceptable distortion has been reached. 830 8. Wavelet Methods 8.16.1 Example This example follows the one in [Shapiro 93]. Figure 8.63a shows three levels of the wavelet transform of an 8Χ8 image. The largest value is 63, so the initial threshold can be anywhere in the range (31, 64]. We set it to 32. Table 8.63b lists the results of the first dominant pass. 63 -34 49 10 7 13 -12 7 -31 23 14 -13 3 4 6 -1 15 14 3 -12 5 -7 3 9 -9 -7 -14 8 4 -2 3 2 -5 9 -1 47 4 6 -2 2 3 0 -3 2 3 -2 0 4 2 -3 6 -4 3 6 3 6 5 11 5 6 0 3 -4 4 Coeff. Reconstr. Subband value Symbol value Note LL3 63 POS 48 1 HL3 -34 NEG -48 LH3 -31 IZ 0 2 HH3 23 ZTR 0 3 HL2 49 POS 48 HL2 10 ZTR 0 4 HL2 14 ZTR 0 HL2 -13 ZTR 0 LH2 15 ZTR 0 LH2 14 IZ 0 5 LH2 -9 ZTR 0 LH2 -7 ZTR 0 HL1 7 Z 0 HL1 13 Z 0 HL1 3 Z 0 HL1 4 Z 0 LH1 -1 Z 0 LH1 47 POS 48 6 LH1 -3 Z 0 LH1 -2 Z 0 (a) (b) Figure 8.63: An EZW Example: Three Levels of an 8Χ8 Image. Notes: 1. The top-left coefficient is 63. It is greater than the threshold, and it is positive, so a POS symbol is generated and is transmitted by the encoder (and the 63 is changed to 0). The decoder assigns this POS symbol the value 48, the midpoint of the interval [32, 64). 2. The coefficient 31 is insignificant with respect to 32, but it is not a zerotree root, since one of its descendants (the 47 in LH1) is significant. The 31 is therefore an isolated zero (IZ). 3. The 23 is less than 32. Also, all its descendants (the 3, -12, -14, and 8 in HH2, and all of HH1) are insignificant. The 23 is therefore a zerotree root (ZTR). As a result, no symbols will be generated by the encoder in the dominant pass for its descendants (this is why none of the HH2 and HH1 coefficients appear in the table). 8.17 DjVu 831 4. The 10 is less than 32, and all its descendants (the -12, 7, 6, and -1 in HL1) are also less than 32. Thus, the 10 becomes a zerotree root (ZTR). Notice that the -12 is greater, in absolute value, than the 10, but is still less than the threshold. 5. The 14 is insignificant with respect to the threshold, but one of its children (they are -1, 47, -3, and 2) is significant. Thus, the 14 becomes an IZ. 6. The 47 in subband LH1 is significant with respect to the threshold, so it is coded as POS. It is then changed to zero, so that a future pass (with a threshold of 16) will code its parent, 14, as a zerotree root. Four significant coefficients were transmitted during the first dominant pass. All that the decoder knows about them is that they are in the interval [32, 64). They will be refined during the first subordinate pass, so the decoder will be able to place them either in [32, 48) (if it receives a 0) or in [48, 64) (if it receives a 1). The encoder generates and transmits the bits “1010” for the four significant coefficients 63, 34, 49, and 47. Thus, the decoder refines them to 56, 40, 56, and 40, respectively. In the second dominant pass, only those coefficients not yet found to be significant are scanned and tested. The ones found significant are treated as zero when the encoder checks for zerotree roots. This second pass ends up identifying the -31 in LH3 as NEG, the 23 in HH3 as POS, the 10, 14, and -3 in LH2 as zerotree roots, and also all four coefficients in LH2 and all four in HH2 as zerotree roots. The second dominant pass stops at this point, since all other coefficients are known to be insignificant from the first dominant pass. The subordinate list contains, at this point, the six magnitudes 63, 49, 34, 47, 31, and 23. They represent the 16-bit-wide intervals [48, 64), [32, 48), and [16, 31). The encoder outputs bits that define a new subinterval for each of the three. At the end of the second subordinate pass, the decoder could have identified the 34 and 47 as being in different intervals, so the six magnitudes are ordered as 63, 49, 47, 34, 31, and 23. The decoder assigns them the refined values 60, 52, 44, 36, 28, and 20. (End of example.) 8.17 DjVu Image compression methods are normally designed for one type of image. JBIG (Section 7.14), for example, was designed for bi-level images, the FABD block decomposition method (Section 7.32) is intended for the compression of discrete-tone images, and JPEG (Section 7.10) works best on continuous-tone images. Certain images, however, combine the properties of all three image types. An important example of such an image is a scanned document containing text, line drawings, and regions with continuous-tone pictures, such as paintings or photographs. Libraries all over the world are currently digitizing their holdings by scanning and storing them in compressed format on disks and CD-ROMs. Organizations interested in making their documents available to the public (such as a mail-order firm with a colorful catalog, or a research institute with scientific papers) also have collections of documents. They can all benefit from an efficient lossy compression method that can highly compress such documents. Viewing such a document is normally done in a web browser, so such a method should feature fast decoding. Such a method is DjVu (pronounced “d΄ej`a vu”), from AT&T laboratories ([ATT 96] and [Haffner et al. 98]). We start with a short summary of its performance. 832 8. Wavelet Methods DjVu routinely achieves compression factors as high as 1000—which is 5 to 10 times better than competing image compression methods. Scanned pages at 300 dpi in full color are typically compressed from 25 Mb down to 30–60 Kb with excellent quality. Black-and-white pages become even smaller, typically compressed to 10–30 Kb. This creates high-quality scanned pages whose size is comparable to that of an average HTML page. For color documents with both text and pictures, DjVu files are typically 5–10 times smaller than JPEG at similar quality. For black-and-white pages, DjVu files are typically 10–20 times smaller than JPEG and five times smaller than GIF. DjVu files are also about five times smaller than PDF files (Section 11.13) produced from scanned documents. To help users read DjVu-compressed documents in a web browser, the developers have implemented a decoder in the form of a plug-in for standard web browsers. With this decoder (freely available for all popular platforms) it is easy to pan the image and zoom on it. The decoder also uses little memory, typically 2 Mbyte of RAM for images that normally require 25 Mbyte of RAM to fully display. The decoder keeps the image in RAM in a compact form, and decompresses, in real time, only the part that is actually displayed on the screen. The DjVu method is progressive. The viewer sees an initial version of the document very quickly, and the visual quality of the display improves progressively as more bits arrive. For example, the text of a typical magazine page appears in just three seconds over a 56 Kbps modem connection. In another second or two, the first versions of the pictures and backgrounds appear. The final, full-quality, version of the page is completed after a few more seconds. We next outline the main features of DjVu. The main idea is that the different elements of a scanned document, namely text, drawings, and pictures, have different perceptual characteristics. Digitized text and line drawings require high spatial resolution but little color resolution. Pictures and backgrounds, on the other hand, can be coded at lower spatial resolution, but with more bits per pixel (for high color resolution). We know from experience that text and line diagrams should be scanned and displayed at 300 dpi or higher resolution, since at any lower resolution text is barely legible and lines lose their sharp edges. Also, text normally uses one color, and drawings use few colors. Pictures, on the other hand, can be scanned and viewed at 100 dpi without much loss of picture detail if adjacent pixels can have similar colors (i.e., if the number of available colors is large). DjVu therefore starts by decomposing the document into three components: mask, foreground, and background. The background component contains the pixels that constitute the pictures and the paper background. The mask contains the text and the lines in bi-level form (i.e., one bit per pixel). The foreground contains the color of the mask pixels. The background is a continuous-tone image and can be compressed at the low resolution of 100 dpi. The foreground normally contains large uniform areas and is also compressed as a continuous-tone image at the same low resolution. The mask is left at 300 dpi but can be efficiently compressed, because it is bi-level. The background and foreground are compressed with a wavelet-based method called IW44 (“IW” stands for “integer wavelet”), while the mask is compressed with JB2, a version of JBIG2 (Section 7.15) developed at AT&T. 8.17 DjVu 833 The decoder decodes the three components, increases the resolution of the background and foreground components back to 300 dpi, and generates each pixel in the final decompressed image according to the mask. If a mask pixel is 0, the corresponding image pixel is taken from the background. If the mask pixel is 1, the corresponding image pixel is generated in the color of the foreground pixel. The rest of this section describes the image separation method used by DjVu. This is a multiscale bicolor clustering algorithm based on the concept of dominant colors. Imagine a document containing just black text on a white background. In order to obtain best results, such a document should be scanned as a grayscale, antialiased image. This results in an image that has mostly black and white pixels, but also gray pixels of various shades located at and near the edges of the text characters. It is obvious that the dominant colors of such an image are black and white. In general, given an image with several colors or shades of gray, its two dominant colors can be identified by the following algorithm: Step 1: Initialize the background color b to white and the foreground color f to black. Step 2: Loop over all the pixels of the image. For each pixel, calculate the distances f and b between the pixel’s color and the current foreground and background colors. Select the shorter of the two distances and flag the pixel as either f or b accordingly. Step 3: Calculate the average color of all the pixels that are flagged f. This becomes the new foreground color. Do the same for the background color. Step 4: Repeat steps 2 and 3 until the two dominant colors converge (i.e., until they vary by less than the value of a preset threshold). This algorithm is simple, and it converges rapidly. However, a real document seldom has two dominant colors, so DjVu uses two extensions of this algorithm. The first extension is called block bicolor clustering. It divides the document into small rectangular blocks, and executes the clustering algorithm above in each block to get two dominant colors. Each block is then separated into a foreground and a background using these two dominant colors. This extension is not completely satisfactory because of the following problems involving the block size: 1. A block should be small enough to have a dominant foreground color. Imagine, for instance, a red word in a region of black text. Ideally, we want such a word to be in a block of its own, with red as one of the dominant colors of the block. If the blocks are large, this word may end up being a small part of a block whose dominant colors will be black and white; the red will effectively disappear. 2. On the other hand, a small block may be located completely outside any text area. Such a block contains just background pixels and should not be separated into background and foreground. A block may also be located completely inside a large character. Such a block is all foreground and should not be separated. In either case, this extension of the clustering algorithm will not find meaningfully dominant colors. 3. Experience shows that in a small block this algorithm does not always select the right colors for foreground and background. The second extension is called multiscale bicolor clustering. This is an iterative algorithm that starts with a grid of large blocks and applies the block bicolor clustering algorithm to each. The result of this first iteration is a pair of dominant colors for each large block. In the second and subsequent iterations, the grid is divided into 834 8. Wavelet Methods smaller and smaller blocks, and each is processed with the original clustering algorithm. However, this algorithm is slightly modified, since it now attracts the new foreground and background colors of an iteration not toward black and white but toward the two dominant colors found in the previous iteration. Here is how it works: 1. For each small block b, identify the larger block B (from the preceding iteration) inside which b is located. Initialize the background and foreground colors of b to those of B. 2. Loop over the entire image. Compare the color of a pixel to the background and foreground colors of its block and tag the pixel according to the smaller of the two distances. 3. For each small block b, calculate a new background color by averaging (1) the colors of all pixels in b that are tagged as background, and (2) the background color that was used for b in this iteration. This is a weighted average where the pixels of (1) are given a weight of 80%, and the background color of (2) is assigned a weight of 20%. The new foreground color is calculated similarly. 4. Steps 2 and 3 are repeated until both background and foreground converge. The multiscale bicolor clustering algorithm then divides each small block b into smaller blocks and repeats. This process is usually successful in separating the image into the right foreground and background components. To improve it even more, it is followed by a variety of filters that are applied to the various foreground areas. They are designed to find and eliminate the most obvious mistakes in the identification of foreground parts. d΄ej`a vu—French for “already seen.” Vuja De—The feeling you’ve never been here. 8.18 WSQ, Fingerprint Compression Most of us don’t realize it, but fingerprints are “big business.” The FBI started collecting fingerprints in the form of inked impressions on paper cards back in 1924, and today they have about 200 million cards, occupying an acre of filing cabinets in the J. Edgar Hoover building in Washington, D.C. (The FBI, like many of us, never throws anything away. They also have many “repeat customers,” which is why “only” about 29 million out of the 200 million cards are distinct; these are the ones used for running background checks.) What’s more, these cards keep accumulating at a rate of 30,000–50,000 new cards per day (this is per day, not per year)! There’s clearly a need to digitize this collection, so it will occupy less space and will lend itself to automatic search and classification. The main problem is size (in bits). When a typical fingerprint card is scanned at 500 dpi, with eight bits/pixel, it results in about 10 Mb of data. Thus, the total size of the digitized collection would be more than 2000 terabytes (a terabyte is 240 bytes); huge even by current (2006) standards. 8.18 WSQ, Fingerprint Compression 835  Exercise 8.18: Apply these numbers to estimate the size of a fingerprint card. Compression is therefore a must. At first, it seems that fingerprint compression must be lossless because of the small but important details involved. However, lossless image compression methods produce typical compression ratios of 0.5, whereas in order to make a serious dent in the huge amount of data in this collection, compressions of about 1 bpp or better are needed. What is needed is a lossy compression method that results in graceful degradation of image details, and does not introduce any artifacts into the reconstructed image. Most lossy image compression methods involve the loss of small details and are therefore unacceptable, since small fingerprint details, such as sweat pores, are admissible points of identification in court. This is where wavelets come into the picture. Lossy wavelet compression, if carefully designed, can satisfy the criteria above and result in efficient compression where important small details are preserved or are at least identifiable. Figure 8.64a,b (obtained, with permission, from Christopher M. Brislawn), shows two examples of fingerprints and one detail, where ridges and sweat pores can clearly be seen. (a) (b) ridge sweat pore Figure 8.64: Examples of Scanned Fingerprints (courtesy Christopher Brislawn). Compression is also necessary, because fingerprint images are routinely sent between law enforcement agencies. Overnight delivery of the actual card is too slow and risky (there are no backup cards), and sending 10 Mb of data through a 9600 baud modem takes about three hours. The method described here [Bradley, Brislawn, and Hopper 93] has been adopted by the FBI as its standard for fingerprint compression [Federal Bureau of Investigations 93]. It involves three steps: (1) a discrete wavelet transform, (2) adaptive scalar quantization of the wavelet transform coefficients, and (3) a two-pass Huffman coding of the quantization indices. This is the reason for the name wavelet/scalar quantization, or WSQ. The method typically produces compression factors of about 20. Decoding is the opposite of encoding, so WSQ is a symmetric compression method. The first step is a symmetric discrete wavelet transform (SWT) using the symmetric filter coefficients listed in Table 8.65 (where R indicates the real part of a complex 836 8. Wavelet Methods number). They are symmetric filters with seven and nine impulse response taps, and they depend on the two numbers x1 (real) and x2 (complex). The final standard adopted by the FBI uses the values x1 = A + B - 1 6, x2 = -(A + B) 2 - 1 6 + i.3(A - B) 2 , where A = -14.15 + 63 1080.15 1/3 , and B = -14.15 - 63 1080.15 1/3 . Tap Exact value Approximate value h0(0) -5.2x1(48|x2|2 - 16Rx2 + 3)/32 0.852698790094000 h0(±1) -5.2x1(8|x2|2 -Rx2)/8 0.377402855612650 h0(±2) -5.2x1(4|x2|2 + 4Rx2 - 1)/16 -0.110624404418420 h0(±3) -5.2x1(Rx2)/8 -0.023849465019380 h0(±4) -5.2x1/64 0.037828455506995 h1(-1) .2(6x1 - 1)/16x1 0.788485616405660 h1(-2, 0) -.2(16x1 - 1)/64x1 -0.418092273222210 h1(-3, 1) .2(2x1 + 1)/32x1 -0.040689417609558 h1(-4, 2) -.2/64x1 0.064538882628938 Table 8.65: Symmetric Wavelet Filter Coefficients for WSQ. The wavelet image decomposition is different from those discussed in Section 8.10. It can be called symmetric and is shown in Figure 8.66. The SWT is first applied to the image rows and columns, resulting in 4Χ4 = 16 subbands. The SWT is then applied in the same manner to three of the 16 subbands, decomposing each into 16 smaller subbands. The last step is to decompose the top-left subband into four smaller ones. The larger subbands (51–63) contain the fine-detail, high-frequency information of the image. They can later be heavily quantized without loss of any important information (i.e., information needed to classify and identify fingerprints). In fact, subbands 60–63 are completely discarded. Subbands 7–18 are important. They contain that portion of the image frequencies that corresponds to the ridges in a fingerprint. This information is important and should be quantized lightly. The transform coefficients in the 64 subbands are floating-point numbers to be denoted by a. They are quantized to a finite number of floating-point numbers that are denoted by ˆa. The WSQ encoder maps a transform coefficient a to a quantization index p (an integer that is later mapped to a code that is itself Huffman encoded). The index p can be considered a pointer to the quantization bin where a lies. The WSQ decoder receives an index p and maps it to a value ˆa that is close, but not identical, to a. This 8.18 WSQ, Fingerprint Compression 837 0 1 2 3 10 16 18 40 42 48 50 51 54 55 56 57 60 61 58 59 62 63 52 53 8 21 27 29 19 22 28 30 20 25 31 33 23 26 32 34 24 9 15 17 39 41 47 49 7 6 12 14 36 38 44 46 5 11 13 35 37 43 45 4 Figure 8.66: Symmetric Image Wavelet Decomposition. is how WSQ loses image information. The set of ˆa values is a discrete set of floatingpoint numbers called the quantized wavelet coefficients. The quantization depends on parameters that may vary from subband to subband, since different subbands have different quantization requirements. Figure 8.67 shows the setup of quantization bins for subband k. Parameter Zk is the width of the zero bin, and parameter Qk is the width of the other bins. Parameter C is in the range [0, 1]. It determines the reconstructed value ˆa. For C = 0.5, for example, the reconstructed value for each quantization bin is the center of the bin. Equation (8.29) shows how parameters Zk and Qk are used by the WSQ encoder to quantize a transform coefficient ak(m, n) (i.e., a coefficient in position (m, n) in subband k) to an index pk(m, n) (an integer), and how the WSQ decoder computes a quantized coefficient ˆak(m, n) from that index: pk(m, n) = ..... .. #ak(m,n)-Zk/2 Qk $+ 1, ak(m, n) > Zk/2, 0, -Zk/2 . ak(m, n) . Zk/2, %ak(m,n)+Zk/2 Qk &+ 1, ak(m, n) < -Zk/2, (8.29) ˆak(m, n) = .. .. . pk(m, n) - CQk + Zk/2, pk(m, n) > 0, 0, pk(m, n) = 0, pk(m, n) + CQk - Zk/2, pk(m, n) < 0. 838 8. Wavelet Methods The final standard adopted by the FBI uses the value C = 0.44 and determines the bin widths Qk and Zk from the variances of the coefficients in the different subbands in the following steps: Step 1: Let the width and height of subband k be denoted by Xk and Yk, respectively. We compute the six quantities Wk = '3Xk 4 (, Hk = '7Yk 16 (, x0k = 'Xk 8 (, x1k = x0k +Wk - 1, y0k = '9Yk 32 (, y1k = y0k + Hk - 1. Step 2: Assuming that position (0, 0) is the top-left corner of the subband, we use the subband region from position (x0k, y0k) to position (x1k, y1k) to estimate the variance .2 k of the subband by .2 k = 1 W·H - 1 x1k n=x0k y1k m=y0kak(m, n) - ΅k2 , where ΅k denotes the mean of ak(m, n) in the region. Step 3: Parameter Qk is computed by qQk =.. .. . 1, 0 . k . 3, 10 Ak loge(.2 k) , 4 . k . 59, and .2 k . 1.01, 0, 60 . k . 63, or .2 k < 1.01, where q is a proportionality constant that controls the bin widths Qk and thereby the overall level of compression. The procedure for computing q is complex and will not be described here. The values of the constants Ak are Ak =...... ... 1.32, k= 52, 56, 1.08, k= 53, 58, 1.42, k= 54, 57, 1.08, k= 55, 59, 1, otherwise. Notice that the bin widths for subbands 60–63 are zero. As a result, these subbands, containing the finest detail coefficients, are simply discarded. Step 4: The width of the zero bin is set to Zk = 1.2Qk. The WSQ encoder computes the quantization indices pk(m, n) as shown, then maps them to the 254 codes shown in Table 8.68. These values are encoded with Huffman codes (using a two-pass process), and the Huffman codes are then written on the compressed stream. A quantization index pk(m, n) can be any integer, but most are small and there are many zeros. Thus, the codes of Table 8.68 are divided into three groups. The first 8.18 WSQ, Fingerprint Compression 839 Z/2 Z/2+Q Z/2+Q(1-C) Z/2+Q(2-C) Z/2+Q(3-C) Z/2+3Q a a Figure 8.67: WSQ Scalar Quantization. group consists of 100 codes (codes 1 through 100) for run lengths of 1 to 100 zero indices. The second group is codes 107 through 254. They specify small indices, in the range [-73, +74]. The third group consists of the six escape codes 101 through 106. They indicate large indices or run lengths of more than 100 zero indices. Code 180 (which corresponds to an index pk(m, n) = 0) is not used, because this case is really a run length of a single zero. An escape code is followed by the (8-bit or 16-bit) raw value of the index (or size of the run length). Here are some examples: An index pk(m, n) = -71 is coded as 109. An index pk(m, n) = -1 is coded as 179. An index pk(m, n) = 75 is coded as 101 (escape for positive 8-bit indices) followed by 75 (in eight bits). An index pk(m, n) = -259 is coded as 104 (escape for negative large indices) followed by 259 (the absolute value of the index, in 16 bits). An isolated index of zero is coded as 1, and a run length of 260 zeros is coded as 106 (escape for large run lengths) followed by 260 (in 16 bits). Indices or run lengths that require more than 16 bits cannot be encoded, but the particular choice of the quantization parameters and the wavelet transform virtually guarantee that large indices will never be generated. O’Day figured that that was more than he’d had the right to expect under the circumstances. A fingerprint identification ordinarily required ten individual points—the irregularities that constituted the art of fingerprint identification—but that number had always been arbitrary. The inspector was certain that Cutter had handled this computer disk, even if a jury might not be completely sure, if that time ever came. —Tom Clancy, Clear and Present Danger The last step is to prepare the Huffman code tables. They depend on the image, so they have to be written on the compressed stream. The standard adopted by the FBI specifies that subbands be grouped into three blocks and all the subbands in a group use the same Huffman code table. This facilitates progressive transmission of the image. The first block consists of the low- and mid-frequency subbands 0–18. The second and third blocks contain the highpass detail subbands 19–51 and 52–59, respectively (recall that subbands 60–63 are completely discarded). Two Huffman code tables are prepared, one for the first block and the other for the second and third blocks. A Huffman code table for a block of subbands is prepared by counting the number of times each of the 254 codes of Table 8.68 appears in the block. The counts are used to determine the length of each code and to construct the Huffman code tree. This is a 840 8. Wavelet Methods Code Index or run length 1 run length of 1 zeros 2 run length of 2 zeros 3 run length of 3 zeros ... 100 run length of 100 zeros 101 escape code for positive 8-bit index 102 escape code for negative 8-bit index 103 escape code for positive 16-bit index 104 escape code for negative 16-bit index 105 escape code for zero run, 8-bit 106 escape code for zero run, 16-bit 107 index value -73 108 index value -72 109 index value -71 ... 179 index value -1 180 unused 181 index value 1 ... 253 index value 73 254 index value 74 Table 8.68: WSQ Codes for Quantization Indices and Run Lengths. two-pass job (one pass to determine the code tables and another pass to encode), and it is done in a way similar to the use of the Huffman code by JPEG (Section 7.10.4). 8.19 JPEG 2000 This section was originally written in mid-2000 and was slightly improved in early 2003. The data compression field is very active, with new approaches, ideas, and techniques being developed and implemented all the time. JPEG (Section 7.10) is widely used for image compression but is not perfect. The use of the DCT on 8Χ8 blocks of pixels results sometimes in a reconstructed image that has a blocky appearance (especially when the JPEG parameters are set for much loss of information). This is why the JPEG committee has decided, as early as 1995, to develop a new, wavelet-based standard for the compression of still images, to be known as JPEG 2000 (or JPEG Y2K). Perhaps the most important milestone in the development of JPEG 2000 occurred in December 1999, when the JPEG committee met in Maui, Hawaii and approved the first committee draft of Part 1 of the JPEG 2000 standard. At its Rochester meeting in August 2000, the JPEG committee approved the final draft of this International Standard. In December 2000 this draft was finally accepted as a full International Standard by the ISO and 8.19 JPEG 2000 841 ITU-T. Following is a list of areas where this new standard is expected to improve on existing methods: High compression efficiency. Bitrates of less than 0.25 bpp are expected for highly detailed grayscale images. The ability to handle large images, up to 232Χ232 pixels (the original JPEG can handle images of up to 216Χ216). Progressive image transmission (Section 7.13). The proposed standard can decompress an image progressively by SNR, resolution, color component, or region of interest. Easy, fast access to various points in the compressed stream. The decoder can pan/zoom the image while decompressing only parts of it. The decoder can rotate and crop the image while decompressing it. Error resilience. Error-correcting codes can be included in the compressed stream, to improve transmission reliability in noisy environments. The main sources of information on JPEG 2000 are [JPEG 00] and [Taubman and Marcellin 02]. This section, however, is based on [ISO/IEC 00], the final committee draft (FCD), released in March 2000. This document defines the compressed stream (referred to as the bitstream) and the operations of the decoder. It contains informative sections about the encoder, but any encoder that produces a valid bitstream is considered a valid JPEG 2000 encoder. One of the new, important approaches to compression introduced by JPEG 2000 is the “compress once, decompress many ways” paradigm. The JPEG 2000 encoder selects a maximum image quality Q and maximum resolution R, and it compresses an image using these parameters. The decoder can decompress the image at any image quality up to and including Q and at any resolution less than or equal to R. Suppose that an image I was compressed into B bits. The decoder can extract A bits from the compressed stream (where A < B) and produce a lossy decompressed image that will be identical to the image obtained if I was originally compressed lossily to A bits. In general, the decoder can decompress the entire image in lower quality and/or lower resolution. It can also decompress parts of the image (regions of interest) at either maximum or lower quality and resolution. Even more, the decoder can extract parts of the compressed stream and assemble them to create a new compressed stream without having to do any decompression. Thus, a lower-resolution and/or lower-quality image can be created without the decoder having to decompress anything. The advantages of this approach are (1) it saves time and space and (2) it prevents the buildup of image noise, common in cases where an image is lossily compressed and decompressed several times. JPEG 2000 also makes it possible to crop and transform the image. When an image is originally compressed, several regions of interest may be specified. The decoder can access the compressed data of any region and write it as a new compressed stream. This necessitates some special processing around the region’s borders, but there is no need to decompress the entire image and recompress the desired region. In addition, mirroring (flipping) the image or rotating it by 90., 180., and 270. can be carried out almost entirely in the compressed stream without decompression. 842 8. Wavelet Methods Image progression is also supported by JPEG 2000 and has four aspects: quality, resolution, spatial location, and component. As the decoder inputs more data from the compressed stream it improves the quality of the displayed image. An image normally becomes recognizable after only 0.05 bits/pixel have been input and processed by the decoder. After 0.25 bits/pixel have been input and processed, the partially decompressed image looks smooth, with only minor compression artifacts. When the decoder starts, it uses the first data bytes input to create a small version (a thumbnail) of the image. As more data is read and processed, the decoder adds more pixels to the displayed image, thereby increasing its size in steps. Each step increases the image by a factor of 2 on each side. This is how resolution is increased progressively. This mode of JPEG 2000 corresponds roughly to the hierarchical mode of JPEG. Spatial location means that the image can be displayed progressively in raster order, row by row. This is useful when the image has to be decompressed and immediately sent to a printer. Component progression has to do with the color components of the image. Most images have either one or three color components, but four components, such as CMYK, are also common. JPEG 2000 allows for up to 16,384 components. Extra components may be overlays containing text and graphics. If the compressed stream is prepared for progression by component, then the image is first decoded as grayscale, then as a color image, then the various overlays are decompressed and displayed to add realism and detail. The four aspects of progression may be mixed within a single compressed stream. A particular stream may be prepared by the encoder to create, for example, the following result. The start of the stream may contain data enough to display a small, grayscale image. The data that follows may add color to this small image, followed by data that increases the resolution several times, followed, in turn, by an increase in quality until the user decides that the quality is high enough to permit printing the image. At that point, the resolution can be increased to match that of the available printer. If the printer is black and white, the color information in the remainder of the compressed image can be skipped. All this can be done while only the data required by the user needs be input and decompressed. Regardless of the original size of the compressed stream, the compression ratio is effectively increased if only part of the compressed stream needs be input and processed. How does JPEG 2000 work? The following paragraph is a short summary of the algorithm. Certain steps are described in more detail in the remainder of this section. If the image being compressed is in color, it is divided into three components. Each component is partitioned into rectangular, nonoverlapping regions called tiles, that are compressed individually. A tile is compressed in four main steps. The first step is to compute a wavelet transform that results in subbands of wavelet coefficients. Two such transforms, an integer and a floating point, are specified by the standard. There are L+1 resolution levels of subbands, where L is a parameter determined by the encoder. In step two, the wavelet coefficients are quantized. This is done if the user specifies a target bitrate. The lower the bitrate, the coarser the wavelet coefficients have to be quantized. Step three uses the MQ coder (an encoder similar to the QM coder, Section 5.11) to arithmetically encode the wavelet coefficients. The EBCOT algorithm [Taubman 99] has been adopted for the encoding step. The principle of EBCOT is to divide each subband 8.19 JPEG 2000 843 into blocks (termed code-blocks) that are coded individually. The bits resulting from coding several code-blocks become a packet and the packets are the components of the bitstream. The last step is to construct the bitstream. This step places the packets, as well as many markers, in the bitstream. The markers can be used by the decoder to skip certain areas of the bitstream and to reach certain points quickly. Using markers, the decoder can, e.g., decode certain code-blocks before others, thereby displaying certain regions of the image before other regions. Another use of the markers is for the decoder to progressively decode the image in one of several ways. The bitstream is organized in layers, where each layer contains higher-resolution image information. Thus, decoding the image layer by layer is a natural way to achieve progressive image transmission and decompression. Before getting to the details, the following is a short history of the development effort of JPEG 2000. The history of JPEG 2000 starts in 1995, when Ricoh Inc. submitted the CREW algorithm (compression with reversible embedded wavelets; Section 8.15) to the ISO/IEC as a candidate for JPEG-LS (Section 7.11). CREW was not selected as the algorithm for JPEG-LS, but was sufficiently advanced to be considered a candidate for the new method then being considered by the ISO/IEC. This method, later to become known as JPEG 2000, was approved as a new, official work item, and a working group (WG1) was set for it in 1996. In March 1997, WG1 called for proposals and started evaluating them. Of the many algorithms submitted, the WTCQ method (wavelet trellis coded quantization) performed the best and was selected in November 1997 as the reference JPEG 2000 algorithm. The WTCQ algorithm includes a wavelet transform and a quantization method. In November 1998, the EBCOT algorithm was presented to the working group by its developer David Taubman and was adopted as the method for encoding the wavelet coefficients. In March 1999, the MQ coder was presented to the working group and was adopted by it as the arithmetic coder to be used in JPEG 2000. During 1999, the format of the bitstream was being developed and tested, with the result that by the end of 1999 all the main components of JPEG 2000 were in place. In December 1999, the working group issued its committee draft (CD), and in March 2000, it has issued its final committee draft (FCD), the document on which this section is based. In August 2000, the JPEG group met and decided to approve the FCD as a full International Standard in December 2000. This standard is now known as ISO/IEC-15444, and its formal description is available (in 13 parts, of which part 7 has been abandoned) from the ISO, ITU-T, and certain national standards organizations. This section continues with details of certain conventions and operations of JPEG 2000. The goal is to illuminate the key concepts in order to give the reader a general understanding of this new international standard. Color Components: A color image consists of three color components. The first step of the JPEG 2000 encoder is to transform the components by means of either a reversible component transform (RCT) or an irreversible component transform (ICT). Each transformed component is then compressed separately. If the image pixels have unsigned values (which is the normal case), then the component transform (either RCT or ICT) is preceded by a DC level shifting. This process translates all pixel values from their original, unsigned interval [0, 2s - 1] (where s is the pixels’ depth) to the signed interval [-2s-1, 2s-1 -1] by subtracting 2s-1 from each 844 8. Wavelet Methods value. For s = 4, e.g., the 24 = 16 possible pixel values are transformed from the interval [0, 15] to the interval [-8, +7] by subtracting 24-1 = 8 from each value. The RCT is a decorrelating transform. It can only be used with the integer wavelet transform (which is reversible). Denoting the pixel values of image component i (after a possible DC level shifting) by Ii(x, y) for i = 0, 1, and 2, the RCT produces new values Yi(x, y) according to Y0(x, y) = 'I0(x, y) + 2I1(x, y) + I2(x, y) 4 (, Y1(x, y) = I2(x, y) - I1(x, y), Y2(x, y) = I0(x, y) - I1(x, y). Notice that the values of components Y1 and Y2 (but not Y0) require one more bit than the original Ii values. The ICT is also a decorrelating transform. It can only be used with the floatingpoint wavelet transform (which is irreversible). The ICT is defined by Y0(x, y) = 0.299I0(x, y) + 0.587I1(x, y) + 0.144I2(x, y), Y1(x, y) = -0.16875I0(x, y) - 0.33126I1(x, y) + 0.5I2(x, y), Y2(x, y) = 0.5I0(x, y) - 0.41869I1(x, y) - 0.08131I2(x, y). If the original image components are red, green, and blue, then the ICT is very similar to the YCbCr color representation (Section 9.2). Tiles: Each (RCT or ICT transformed) color component of the image is partitioned into rectangular, nonoverlapping tiles. Since the color components may have different resolutions, they may use different tile sizes. Tiles may have any size, up to the size of the entire image (i.e., one tile). All the tiles of a given color component have the same size, except those at the edges. Each tile is compressed individually. Figure 8.69a shows an example of image tiling. JPEG 2000 allows the image to have a vertical offset at the top and a horizontal offset on the left (the offsets can be zero). The origin of the tile grid (the top-left corner) can be located anywhere inside the intersection area of the two offsets. All the tiles in the grid are the same size, but those located on the edges, such as T0, T4, and T19 in the figure, have to be truncated. Tiles are numbered in raster order. The main reason for having tiles is to enable the user to decode parts of the image (regions of interest). The decoder can identify each tile in the bitstream and decompress just those pixels included in the tile. Figure 8.69b shows an image with an aspect ratio (height/width) of 16 : 9 (the aspect ratio of high-definition television, HDTV; Section 9.3.1). The image is tiled with four tiles whose aspect ratio is 4 : 3 (the aspect ratio of current, analog television), such that tile T0 covers the central area of the image. This makes it easy to crop the image from the original aspect ratio to the 4:3 ratio. Wavelet Transform: Two wavelet transforms are specified by the standard. They are the (9,7) floating-point wavelet (irreversible) and the (5,3) integer wavelet (reversible). Either transform allows for progressive transmission, but only the integer transform can produce lossless compression. 8.19 JPEG 2000 845 (a) T0 T0 T1 T2 T3 T1 T2 T3 T4 T5 T15 T10 T6 T9 T14 T18 T19 (b) Vertical offset Vertical offset Horizontal offset Horizontal offset Image Tile grid Figure 8.69: JPEG 2000 Image Tiling. We denote a row of pixels in a tile by Pk, Pk+1,. . . , Pm. Because of the nature of the wavelet transforms used by JPEG 2000, a few pixels with indices less than k or greater than m may have to be used. Therefore, before any wavelet transform is computed for a tile, the pixels of the tile may have to be extended. The JPEG 2000 standard specifies a simple extension method termed periodic symmetric extension, which is illustrated by Figure 8.70. The figure shows a row of seven symbols “ABCDEFG” and how they are reflected to the left and to the right to extend the row by l and r symbols, respectively. Table 8.71 lists the minimum values of l and r as functions of the parity of k and m and of the particular transform used. l ‹------------- r -------------› . . .DEFGFEDCB|ABCDEFG|FEDCBABCD. . . ^k ^m Figure 8.70: Extending a Row of Pixels. k l(5,3) l (9,7) m r(5,3) r (9,7) even 2 4 odd 2 4 odd 1 3 even 1 3 Table 8.71: Minimum Left and Right Extensions. We now denote a row of extended pixels in a tile by Pk, Pk+1,. . . , Pm. Since the pixels have been extended, index values below k and above m can be used. The (5,3) integer wavelet transform computes wavelet coefficients C(i) by first computing the odd values C(2i + 1) and then using them to compute the even values C(2i). The 846 8. Wavelet Methods computations are C(2i + 1) = P(2i + 1) - 'P(2i) + P(2i + 2) 2 (, for k - 1 . 2i + 1> 4 instead of the equivalent real expression 0.375a + 0.125b + 0.125c + 0.375d. bil[u_,w_]:=a(1-u)(1-w)+b(1-u)w+d(1-w)u+c u w; bil[.25, .25] bil[.75, .5] bil[.25, 1] bil[.5, .25] 0.5625a+0.1875b+0.0625c+0.1875d 0.125a+0.125b+0.375c+0.375d 0.75b+0.25c 0.375a+0.125b+0.125c+0.375d Figure 9.64: Code For Bilinear Interpolation. The principles of bicubic interpolation (we are interested in the one-dimensional case) are discussed in Section 7.25.5. The idea is to start with the values of four consecutive pixels, consider them the y coordinates of four equally-spaced two-dimensional points p1 through p4, and construct a polynomial P(t) that passes through the points when t varies from 0 to 1. This polynomial is given by Equation (7.45), duplicated here G(t) = (t3, t2, t, 1)... -4.5 13.5 -13.5 4.5 9.0 -22.5 18 -4.5 -5.5 9.0 -4.5 1.0 1.0 0 0 0 ... ... p1 p2 p3 p4 ... . (7.45) The four original points are obtained for t values 0, 1/3, 2/3, and 1 and Figure 9.62d makes it clear that we are interested in the three pixels that are equally spaced between 940 9. Video Compression p2 and p3. These correspond to t values 0.44444, 0.5, and 0.55555, respectively. The Mathematica code of Figure 9.65 shows how this computation is carried out. In order to speed up the decoder, the definition of VC-1 employs integers and a right shift instead of the real numbers listed in the figure. g[t_]:={t^3, t^2, t, 1}. {{-4.5,13.5,-13.5,4.5},{9.,-22.5,18.,-4.5}, {-5.5,9.,-4.5,1.},{1.,0,0,0}}.{p1,p2,p3,p4}; g[0.44444] g[0.5] g[0.55555] -0.0617277p1+0.740754p2+0.370355p3-0.0493812 p4 -0.0625p1+0.5625p2+0.5625p3-0.0625p4 -0.0493846p1+0.37039p2+0.740724p3-0.0617293p4 Figure 9.65: Code For Bicubic Interpolation. Section 7.25.7 shows how cubic interpolation can be extended to two dimensions, but VC-1 specifies this process differently. The VC-1 encoder computes the bicubic interpolation of half- and quarter pixels in several steps. First, this type of interpolation is performed horizontally to compute the three pixels 1, 2, 3 between b and c (Figure 9.62e), using, in addition to b and c, the values of the pixel to the left of b (as our p1) and the pixel to the right of c (as our p4). Next, the three pixels 19, 20, and 21 between a and d are computed in a similar way. Using the values of 1 (and the corresponding pixel in the 5 Χ 5 grid above) and 19 (and the corresponding pixel in the grid below), it is possible to interpolate for pixels 5, 10, and 15. The values of 6, 11, and 16, and 7, 12, and 17 are similarly computed. Finally, the values of 4, 9, and 14 are computed from pixels b (and the one above it) and a (and the one below it), and similarly for 8, 13, and 18. Another tough job for the VC-1 implementor! 9.11.8 Deblocking Filters The VC-1 encoder transforms blocks of pixels and it has long been known that this introduces boundary artifacts into the decompressed frame. Imagine two adjacent identical pixels with a value of, say, 100. Assume further that a block boundary passes between the pixels. After being transformed and quantized by the encoder and later reversetransformed and dequantized by the decoder, one pixel may end up as 102, while the other may become 98. The difference between an original and a reconstructed pixel is 2, but the difference between the two reconstructed pixels is double that. Blocking artifact can also be caused by a poor quantization of the DC coefficient, causing the intensity of a decoded block to become lower or higher than the intensities of its neighbors. VC-1 introduces two innovative techniques to reduce the effect of blocking artifacts. They are referred to as overlapped transform (OLT) and in-loop deblocking filter (ILF). OLT acts as a smoothing filter. It is designed to perform best when two adjacent 8Χ8 blocks differ much in their quality. Imagine a block P on the left and a block Q on the right with a vertical boundary between them (alternatively, P may be directly above 9.11 VC-1 941 Q, with a horizontal boundary between them). If P and Q differ in their qualities (one is high quality and the other is low quality), then the encoder applies the OLT filter to smooth their boundary. This operation is carried out only on intra-coded blocks, which are always 8 Χ 8. The formal definition places many restrictions on the OLT filter, so the encoder has to signal to the decoder the presence of this operation. Once the encoder decides to apply OLT to two adjacent blocks (the codec definition, as usual, does not specify how the encoder should make this decision), it starts with the four edge pixels P1, P0, Q0, and Q1 shown in Figure 9.66 (for two blocks with a horizontal boundary, the figure should be rotated 90.). For each of the eight rows of the blocks, the encoder computes new values for these pixels, based on their original values and on certain rounding parameters, and substitutes the new values for the old ones. The decoder performs the reverse operations. Specifically, the computation carried out by the decoder is ... P1 P0 Q0 Q1 ... =.. .. . ... 7 0 0 1 -1 7 1 1 1 1 7 -1 1 0 0 7 ... ... P1 P0 Q0 Q1 ... +... r0 r1 r0 r1 ... .. .. . >> 3, where the notation >> means a right shift and the two rounding parameters r0 and r1 take the values (3, 4) for even-numbered rows of pixels and (4, 3) for odd-numbered rows.             Figure 9.66: Edge Pixels For OLT. When two adjacent blocks have similar qualities, the VC-1 standard recommends the use of ILF. Notice that an in-loop filter improves prediction since it filters frames that will be used for reference (this is different than a post-filter). For blocks in I and B frames, the ILF filter should be applied to all four boundaries of the 8 Χ 8 blocks in the following way. First, ILF should be applied to every 8th and 9th horizontal rows of the frame, and then to every 8th and 9th vertical columns. If the rows are numbered 0 through N - 1, then ILF should be applied to rows (7, 8), (15, 16),. . . , (8(N -1)-1, 8(N -1), and similarly for the columns (Figure 9.67). Those rows and columns form the block boundaries in the frame. For blocks in P frames, the situation is more complex, because blocks can be of four different sizes, as illustrated by Figure 9.69a. The encoder therefore has to work harder to identify the pixels that constitute block boundaries, as shown in part (b) of 942 9. Video Compression Figure 9.67: ILF Boundary Rows and Columns For I and B Frames. the figure. The figure shows examples where filtering between neighboring blocks (or subblocks) does or does not occur. If the motion vectors of adjacent blocks are different, then the boundary between them is always filtered. Thus, our example assumes that the motion vectors of the blocks are identical. The shaded blocks or subblocks in the figure are those that have at least one nonzero coefficient, while the clear blocks or subblocks have no transform coefficients. The thick lines designate boundaries that should be filtered. The order of filtering is important and is specified in detail in the VC-1 definition. The figure illustrates only horizontal block neighbors. The handling of vertical block neighbors is similar. Once the encoder identifies the block boundaries that should be filtered, the filtering operation can proceed. Figure 9.68 illustrates this process. Filtering is always done on a segment of 2Χ4 (horizontal) or 4Χ2 (vertical) pixels. Thus, if a boundary is eight pixels long, its filtering becomes a two-step process. In each four-pixel segment, the third pair of pixels is filtered first and the result of this operation determines whether the other three pairs should also be filtered.               Figure 9.68: ILF Filtering. The rest of the filtering operation differs slightly for P frames and for the other frame types. It is described here in essence but not in detail. The third pair and its six near neighbors are split into three groups as shown in the figure and several expressions involving differences of pairs of pixels in each group are computed. The idea is to discover any blocking effects caused by differences between pixels on both sides of the boundary. These differences are examined to see whether the remaining three pairs should also 9.11 VC-1 943                          Figure 9.69: ILF Boundary Rows and Columns For P Frames. be filtered. The two pixels p4 and p5 that are closest to the boundary are updated according to d = (p4 - p5)/2, p4 = p4 - d, p5 = p5 + d. If the decision is made to filter the other three pairs, they are updated in a similar way. In each pair, pixel differences are computed and the two pixels closest to the boundary are updated. 9.11.9 The Transform The transform is a central component of VC-1. It is an integer transform that is based on the discrete cosine transform (DCT). Clearly, an integer transform is easier to implement and is faster to execute. Its precision, however, depends on the size of the integers in question, but each color component of a pixel is typically only eight bits long, so a 16- bit transform provides enough precision to generate intermediate results that are much bigger than an original pixel. It has already been mentioned that the transform operates on four different block sizes, as shown in Figure 9.56. This is one of the many innovations of VC-1 and it is an option of the encoder (that is signaled at the sequence layer). Blocks are first motion compensated and then transformed. The intra blocks (blocks in macroblocks in I frames) always use the 8 Χ 8 transform, while the inter blocks (in the other types of frames) may use any of the four transform sizes (and this information is stored in the picture, macroblock, or block layers). When the transform size appears in the picture 944 9. Video Compression layer, all the blocks in the picture use the same transform size. When the size appear at the macroblock layer, all the blocks of the macroblock use the same transform size. The details of this transform are described here, based on [Srinivasan and Regunathan 05], but the interested reader should first become familiar with the concept of orthogonal transforms (Section 7.7) and especially with the matrix of Equation (7.6) and with the DCT matrix in one dimension for n = 8, Table 7.38, shown here as matrix (9.5) (unnormalized) and (9.6) (normalized). Notice that the odd-numbered rows of this matrix have only three distinct elements and the even-numbered rows consist of another set of four distinct elements. ........... 1. 1. 1. 1. 1. 1. 1. 1. 0.981 0.831 0.556 0.195 -0.195 -0.556 -0.831 -0.981 0.924 0.383 -0.383 -0.924 -0.924 -0.383 0.383 0.924 0.831 -0.195 -0.981 -0.556 0.556 0.981 0.195 -0.831 0.707 -0.707 -0.707 0.707 0.707 -0.707 -0.707 0.707 0.556 -0.981 0.195 0.831 -0.831 -0.195 0.981 -0.556 0.383 -0.924 0.924 -0.383 -0.383 0.924 -0.924 0.383 0.195 -0.556 0.831 -0.981 0.981 -0.831 0.556 -0.195 ........... . (9.5) ....... 0.3536 0.3536 0.3536 0.3536 0.3536 0.3536 0.3536 0.3536 0.4904 0.4157 0.2778 0.0975 -0.0975 -0.2778 -0.4157 -0.4904 0.4619 0.1913 -0.1913 -0.4619 -0.4619 -0.1913 0.1913 0.4619 0.4157 -0.0975 -0.4904 -0.2778 0.2778 0.4904 0.0975 -0.4157 0.3536 -0.3536 -0.3536 0.3536 0.3536 -0.3536 -0.3536 0.3536 0.2778 -0.4904 0.0975 0.4157 -0.4157 -0.0975 0.4904 -0.2778 0.1913 -0.4619 0.4619 -0.1913 -0.1913 0.4619 -0.4619 0.1913 0.0975 -0.2778 0.4157 -0.4904 0.4904 -0.4157 0.2778 -0.0975 ....... . (9.6) The first thing that comes to mind, when trying to derive an integer version of the DCT, is to round off the elements of Equation (9.5) to the nearest integer, but it is easy to see that this would simply result in a matrix of zeros and 1’s, similar to the basic orthogonal transform matrix of Equation (7.6). It would still be an orthogonal transform, but without the useful energy compaction properties of the DCT (however, note that energy compaction is not a critical issue here, because the DCT operates on the residuals of the prediction). Thus, a different approach is needed. The designers of the VC-1 transform started with the following considerations: Video encoding can sometimes be slow, but decoding must be done at real time. Video decoding is often done by small, specialized processors that are embedded in DVD players, digital televisions, home theaters, handheld mp3 and video devices, and other devices and instruments designed for video playback. Currently, most such devices are based on 16-bit architectures. It is known from long experience with video compression that the inverse transform is generally responsible for 10–30% of the CPU time used by a video decoder. A single video frame of 1920 Χ 1080 pixel (the so-called HDTV 1080p) consists of 32,400 8 Χ 8 blocks that have to be inverse transformed. At 30 frames per second, the decoder has to inverse-transform 972,000 blocks each second; a tough job for even fast processors. 9.11 VC-1 945 These two considerations imply that a 16-bit integer transform can be very useful if it also offers energy compaction equivalent to that of the DCT. The integer transform developed for VC-1 allows for fast decoding. The decoder performs the inverse transform using just 16-bit integer arithmetic. The encoder, on the other hand, can compute the forward transform with fixed-point (scaled integer) or even floating-point arithmetic. (The encoder can also benefit from a simple transform, because it has to try all the valid block sizes, such as 8 Χ 8 and 4 Χ 8, in order to choose the best.) A full understanding of this transform starts with the concept of norm. In mathematics, the term “norm” is defined for certain mathematical objects that are more complex than a single number. Intuitively, the norm of an object is a quantity that describes the length, magnitude, or extent of the object. Norms are defined for complex numbers, quaternions, vectors, matrices, and other objects that are not just numbers. Often, several different norms are defined for the same type of object. The norm of a matrix can be defined in several ways, but perhaps the most common definition is the Frobenius norm (older texts refer to it as the Hilbert-Schmidt norm or the Schur norm). Given an n Χm matrix A with elements aij , its Frobenius norm is ||A|| =.. n i=1 m j=1 |aij |2.. 1/2 . Similarly, the norm of a vector is the square root of the sum of the squares of its elements. These norms are very similar to the concept of energy, defined in Section 7.6. Given a vector V = (2, 2, 2) and a 3 Χ 3 matrix A all whose elements are 10, the result of the product V · A is the vector W = (60, 60, 60), and it is easy to verify that ||W|| = ||V ||·||A|| = 60.3. We therefore conclude that multiplying a vector by a matrix A scales its norm by (approximately) a factor of ||A||. The two-dimensional DCT (Equation (7.16), duplicated here) can be written in matrix notation as the product of a row of m functions cos[(2y + 1)j./2m], an n Χ m matrix of pixels pxy, and a column of n functions cos[(2x + 1)i./2n]. Thus Gij =  2 m2 n CiCj n-1 x=0 m-1 y=0 pxy cos (2y + 1)j. 2m cos (2x + 1)i. 2n  = row of cos (2y + 1)j. 2m ... p00 . . . pnm ... .... column of cos (2x+1)i. 2n  .... , (9.7) Each coefficient Gij is therefore the product of a row of cosines, a block of pixels, and a column of cosines. An entire block of transform coefficients Gij is therefore the product of the one-dimensional DCT matrix (Equation (7.14)), the corresponding block of pixels, and the transpose of the same DCT matrix. Our task is to replace the DCT matrix with a matrix T of integers, and the earlier discussion of norms makes it clear that the resulting transform coefficients Gij will be bigger than the original pixels by a factor of approximately ||T||2. 946 9. Video Compression We denote L def = log2 ||T||. We assume that the prediction residuals are 9-bit signed integers, i.e., in the interval [-256, 255]. We require that the transform coefficients be 12-bit signed integers. Transforming a block of pixels therefore increases the magnitude of a pixel by three bits. Because transforming is done by two matrix multiplications (by the transform matrix T and its transpose), we conclude that each matrix multiplication increases the magnitude of pixels by 1.5 bits on average. To a large extent, decoding is the reverse of encoding, but the integer nature of the transform creates a problem. Decoding starts with 12-bit signed transform coefficients and has to end up with 9-bit signed integer pixels, but decoding is done by two matrix multiplications which increase the size of the numbers involve. Thus, each matrix multiplication should be followed by a right shift (denoted by >>) which shrinks the results from 12 bits to 10.5 bits and then to nine bits. Assuming that decoding starts with a matrix A of 12-bit transform coefficients, the first step is to multiply B = T ·A and shift C = B >> n, where the value of n should be such that the elements of C will be 10.5 bits long on average. The second step is another multiplication D = C · T followed by another right shift E = D >> m, where m should be chosen such that the elements of E are 9-bit integers. This sequence of transformations and magnitude changes is achieved if we require that ||T||2 = 2n+m, which implies 2L = n + m. Also, the dynamic range (the range of values) of matrix B should be 10.5+L and the range of D should be 9+2L-n. These conditions should be achieved within the constraints of 16-bit arithmetic, which results in the following relations 2L = n + m, 10.5 + L . 16, 9 + 2L - n . 16. One solution of these relations is L . 5.5 (which implies ||T||2 = 211 = 2,048) and m = 6. These values are enough to place a limit on the DC coefficient of the transform. Equation (9.7) implies that the DC coefficient G00 is obtained by multiplying the pixels by the top row of the transform matrix, a row whose elements are identical. If we denote the (yet unknown) element on the top row of our 8 Χ 8 integer transform matrix T by d8, then the norm of that row is 8d28 . This norm must be less than or equal the norm of T, so it must satisfy 8d28 . 2,048 or d8 . 16. For the 4 Χ 4 transform matrix, where each row is half the size of the 8Χ8 matrix, the element of the top row is denoted by d4 and the norm requirement is d4 = .2d8. It is now easy to check every pair of integers (d4, d8) for the constraints d8 . 16 and d4 = .2d8, and a quick search verifies that the only pairs that come within 5% of the constraints are (7, 5), (10, 7), (17, 12) and (20, 14). The next task is to determine the remaining seven rows of the 8Χ8 integer transform matrix. We first examine the even-numbered rows and it is easy to see that they contain only four distinct elements that we denote by C1 through C4. These rows form the 4Χ8 matrix ... C1 C2 C3 C4 -C4 -C3 -C2 -C1 C2 -C4 -C1 -C3 C3 C1 C4 -C2 C3 -C1 C4 C2 -C2 -C4 C1 -C3 C4 -C3 C2 -C1 C1 -C2 C3 -C4 ... . (9.8) The following conditions were set by experts in order to determine the four elements 9.11 VC-1 947 (the curious reader might want to glance at Section 8.7.1, for a list of similar conditions used to determine the values of wavelet filter coefficients): 1. The rows of T must be orthogonal. We know that each even-numbered row is orthogonal to the odd-numbered rows, because of the properties of the cosine function. Thus, we have to choose the four elements such that the four even-numbered rows will be orthogonal. 2. The vector (C1,C2,C3,C4) should correlate well with the special DCT vector (0.4904, 0.4157, 0.2778, 0.0975). The term “correlate” refers to the cosine of the angle between the vectors, as measured by their dot product. The DCT vector above was obtained in the following way. The DCT matrix of Equation (9.6) is normalized. Normalization is done by dividing each row by its length, so it becomes a vector of length 1. When this is done to the second row, it becomes the unit vector (0.4904, 0.4157, 0.2778, 0.0975,-0.0975,-0.2778,-0.4157,-0.4904) and the DCT vector above is the first half of this (its length is therefore 0.5). 3. The norm C12 +C22 +C33 +C42 of the four elements should match that of d8, i.e., it should be as close as possible to d28 + d28 + d28 + d28 = 4d28 . The four choices of d8 listed earlier yield the four values 100, 196, 576, and 784 for 4d28 . How can we derive the set of four integers that come closest to satisfying the three conditions above? The simplest approach is to use brute force (i.e., the high speed of our computers). We loop over all the possible sets of four positive integers and select the sets that satisfy C12 + C22 + C33 + C42 . 800. (There is a very large number of such sets, but it is finite.) For each set, we check to see if the four rows of matrix (9.8) are orthogonal. If so, we compute the dot product of condition 2 above and save it. After all possible sets of four integers have been examined in this way, we select the set whose dot product is the maximal. This turns out to be the set (16, 15, 9, 4), whose norm is 578 (implying that d4 = 17 and d8 = 12), very close to the ideal 576. The correlation of condition 2 is 0.9984, also very encouraging. We next examine the odd-numbered rows of Equation (9.6). We already know that d8 = 12, so these four rows become the 4 Χ 8 matrix ... 12 12 12 12 12 12 12 12 C6 C6 -C6 -C5 -C5 -C6 C6 C5 12 -12 -12 12 12 -12 -12 12 C6 -C5 C5 -C6 -C6 C5 -C5 C6 ... and require the derivation of only two elements, denoted by C5 and C6. The norm C52 +C62 of these elements should match (similar to condition 3 above) that of d4, i.e., it should be as close as possible to d28 + d28 = 288, and this is satisfied, to within 1%, by the pair (16, 6). 948 9. Video Compression The 8 Χ 8 transform matrix T8 has now been fully determined and is given by T8 = ........... 12 12 12 12 12 12 12 12 16 15 9 4 -4 -9 -15 -16 16 6 -6 -16 -16 -6 6 16 15 -4 -16 -9 9 16 4 -15 12 -12 -12 12 12 -12 -12 12 9 -16 4 15 -15 -4 16 -9 6 -16 16 -6 -6 16 -16 6 4 -9 15 -16 16 -15 9 -4 ........... . The 4 Χ 4 integer transform is derived in a similar way, but is much simpler. We already know that the transform matrix should look like the 4Χ4 matrix of Equation (7.6) and that d4 equals 17. Thus, we are looking for a matrix of the form ... 17 17 17 17 D1 D2 -D2 -D1 17 -17 -17 17 D2 -D1 D1 -D2 ... . This matrix is inherently orthogonal for any choice of elements D1 and D2, but the best choices are those that satisfy the norm condition D12+D22 = d24 +d24 = 578, and a direct search yields the pair D1 = 22 and D2 = 10, that satisfy D12 +D22 = 222 +102 = 584. Thus, the complete 4 Χ 4 integer transform used by VC-1 is T4 =... 17 17 17 17 22 10 -10 -22 17 -17 -17 17 10 -22 22 -10 ... . Once the two integer transform matrices have been constructed, we are ready for the transform itself. As has already been mentioned, the forward transform is computed by the encoder, which doesn’t always have to work at real time. Thus, the forward transform can be performed with fixed-point (scaled integer) or even floating-point arithmetic. We start with two normalization vectors c4 = 8 289, 8 292, 8 289, 8 292 , c8 = 8 288, 8 289, 8 292, 8 289, 8 288, 8 289, 8 292, 8 289 . These are used to construct the four normalization matrices N4Χ4, N8Χ4, N4Χ8, and N8Χ8 by means of outer products (also known as tensor products) of the vectors c4 and c8. Thus, for example, N8Χ4 = c8 . c4. The outer product is a vector operation whereby a matrix is constructed from the elements of two vectors. Given the two vectors a = (a1, a2, . . . , am) and b = (b1, b2, . . . , bn), 9.11 VC-1 949 their outer product a . b is the mΧ n matrix .... a1b1 a1b2 . . . a1bn a2b1 a2b2 . . . a2bn . . . amb1 amb2 . . . ambn .... . In contrast, the more familiar inner product (or dot product) operates on the elements of two equal-length vectors to produce a single number. The forward transform starts with a block D of pixels and produces a block ˆD of transform coefficients. The transform is defined by ˆD = (TiDT j) . NiΧj where i and j take the values 4 or 8, and the notation “.” means an entry-wise multiplication of two matrices. Thus, depending on the transform block size, there are the following four cases: ˆD = (T4DT4 ) .N4Χ4, ˆD = (T8DT4 ) .N8Χ4, ˆD = (T4DT8 ) .N4Χ8, ˆD = (T8DT8 ) .N8Χ8. The inverse transform starts with a block of 12-bit signed (i.e., in the interval [-2,048, +2,047]), quantized transform coefficients DiΧj (where i and j are 4 or 8) and results in a block RiΧj of 10-bit signed integers (i.e., in the interval [-512, 511]) that become the reconstructed pixels. The transform consists of the two steps EiΧj = (DiΧj · Tj + 4) >> 3, RiΧj = (T i · EiΧj + Ci . 1j + 64) >> 7, where EiΧj are intermediate matrices whose entries are 13-bit signed integers (i.e., in the interval [-4,096, +4,095]), 1i is a row of i 1’s, Cj are the columns C4 = (0, 0, 0, 0) and C8 = (0, 0, 0, 0, 1, 1, 1, 1), and the notation “>>” means an arithmetic (sign-preserving) right shift, performed on all the entries of a matrix. The boldface 4 and 64 refer to adding these constants to all the elements of a matrix. 9.11.10 Quantization and Scan VC-1 uses both uniform and non-uniform quantization schemes. The uniform quantization method treats the DC coefficients of intra blocks differently from the other transform coefficients. These DC coefficients are quantized by dividing them by a parameter DCStepSize, that is computed as follows DCStepSize =... MQUANT = 1, 2, 2 · MQUANT, MQUANT = 3, 4, 8, MQUANT . 5, 6 + MQUANT/2, where MQUANT is a parameter derived in a complex way from the syntax elements DQBILEVEL, MQDIFF, PQUANT (the latter parameter is the picture quantizer scale parameter, which itself is derived in a complex way from the Picture Quantizer Index, PQINDEX), ALTPQUANT and ABSMQ. All other transform coefficients are quantized by dividing them by 2 · MQUANT (except in some cases, not listed here, where the quantization parameter is 2 · MQUANT + HALFQP). 950 9. Video Compression The non-uniform quantization method also treats the DC coefficients of intra blocks differently from the other transform coefficients. This coefficient is quantized as in the uniform method, while all other transform coefficients are quantized by dividing them by first subtracting MQUANT and then dividing by 2·MQUANT (except in some cases, not listed here, where the transform parameter is divided by 2 ·MQUANT+HALFQP). The decoder performs the reverse operations, then rounds the results to the nearest integer and, if necessary, also truncates them to 12 bits, signed, so they lie in the interval [-2,048, +2,047]. Once the transform coefficients have been quantized, they are collected in a zigzag scan of the block. As most other features of VC-1, the zigzag scan depends on syntax parameters and can be done in different ways. Intra-coded blocks are scanned in normal, horizontal, or vertical mode depending on parameter ACPRED and on the DC prediction direction (which can be top or left) as follows. If ACPRED = 0, perform normal scan. If ACPRED = 1 and prediction direction is top, perform horizontal scan. If ACPRED = 1 and prediction direction is left, perform vertical scan. Figure 9.70 shows the normal and horizontal zigzag scans and cell numbering within an 8 Χ 8 block. 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 Normal scan: 0, 8, 1, 2, 9, 16, 24, 17, 10, 3, 4, 11, 18, 25, 32, 40, 33, 48, 26, 19, 12, 5, 6, 13, 20, 27, 34, 41, 56, 49, 57, 42, 35, 28, 21, 14, 7, 15, 22, 29, 36, 43, 50, 58, 51, 59, 44, 37, 30, 23, 31, 38, 45, 52, 60, 53, 61, 46, 39, 47, 54, 62, 55, 63. Horizontal scan: 0, 1, 8, 2, 3, 9, 16, 24, 17, 10, 4, 5, 11, 18, 25, 32, 40, 48, 33, 26, 19, 12, 6, 7, 31, 20, 27, 34, 41, 56, 49, 57, 42, 35, 28, 21, 14, 15, 22, 29, 36, 43, 50, 58, 51, 44, 37, 30, 23, 31, 38, 45, 52, 59, 60, 53, 46, 39, 47, 54, 61, 62, 55, 63. Figure 9.70: VC-1 Normal and Horizontal Zigzag Scans. For inter-coded blocks, there are four block sizes and VC-1 defines two scan patterns for each, one for the simple and main profiles, and the other for the advanced profile; an implementor’s nightmare! 9.11.11 Entropy Coding The step following quantization is the entropy coding of the quantized transform coefficients. This is also very complex because many different tables of indexes and of variable-length codes are used, depending on the profile/level and the type of block being coded (intra or inter). Instead of listing the many rules, syntax parameters, and tables, we show a typical example that illustrates the basic process. Figure 9.71 shows an 8Χ8 block of quantized transform coefficients, where only eight of the 64 coefficients are nonzero. The block is scanned in basic zigzag pattern that brings the nonzero coefficients to the front, creates several short runs of zeros between them, and ends with a long run of zeros. Each nonzero coefficient C is first converted to a triplet (run, level, flag), where run is the length of the run of zeros preceding C, level is the (signed) value of C, and flag is a single bit, that is zero except for the last triplet. Thus, the eight triplets for our example are (1, 1, 0), (0,-2, 0), (2, 5, 0), (1,-1, 0), (1, 1, 0), (0,-3, 0), (0, 4, 0), and (1, 1, 1). 9.11 VC-1 951                                                                                 Figure 9.71: Example of Variable-Length Coding of a Block. VC-1 employs 65 tables of variable-length codes. In the formal specification of this codec [SMPTE-421M 08], these tables are numbered from 168 to 232 and most are used to encode transform coefficients. Tables 175–190 and 219–232 contain variable-length codes for high-level inter and intra-coded blocks. Tables 191 through 204 are similarly used to encode low-level transform coefficients, and tables 205 through 218 specify codes for mid-level. Each set consists of tables with indexes and tables with the actual codes. For our example, we use the tables for low-motion, specifically table 199 for indexes of triplets of the form (run, level, 0), table 200 for an index for the last triplet, whose format is (run, level, 1), and table 198 for the actual variable-length codes. Table 9.72 lists parts of these three code tables. The first triplet is (1, 1, 0). We search table 199 for run 1 and level 1, and find index 14. Entry 14 of table 198 specifies the 4-bit code 1110 = 1011. Thus, the encoded stream starts with 1011|0, where the single zero implies a positive nonzero coefficient. The second triplet is (0,-2, 0). We search table 199 for run 0 and level 2, and find index 1. Entry 1 of table 198 specifies the 5-bit code 2010 = 10100. The encoded stream becomes 1011|0|10100|1, where the last 1 implies a negative coefficient. The third triplet is (2, 5, 0). We search table 199 for run 2 and level 5, and find index 27. Entry 27 of table 198 specifies the 14-bit code 152210 = 00010111110010. The encoded stream becomes 1011|0|10100|1|00010111110010|0. This process continues until the last triplet (1, 1, 1). Since the flag is 1 (indicating the last triplet), we search table 200 for an index. The table yields index 86 for run 1 and level 1, so we look at entry 86 of table 198 and find the 4-bit code 5 = 0101. The complete encoded bitstream is therefore 1011|0|10100|1|00010111110010|0|1011|1|1011|0|0010111|1|01111111|0|0101|0, and this encodes eight 12-bit signed nonzero quantized transform coefficients. 952 9. Video Compression VLC Index Code Size 0 4 3 1 20 5 2 23 7 3 127 8 4 340 9 5 498 10 . . . . . . 14 11 4 . . . . . . 27 1522 14 . . . . . . 86 5 4 . . . . . . 147 5524 13 Table 198 Index Run Level 0 0 1 1 0 2 2 0 3 3 0 4 4 0 5 5 0 6 . . . . . . 14 1 1 . . . . . . 27 2 5 . . . . . . 80 29 1 Table 199 Index Run Level 81 0 1 82 0 2 83 0 3 84 0 4 85 0 5 86 1 1 . . . . . . 103 7 2 Table 200 Table 9.72: Tables of Indexes and Variable-Length Codes. In seven to ten years video traffic on the Internet will exceed data and voice traffic combined. —Bob Metcalfe 10 Audio Compression Text does not occupy much space in the computer. An average book, consisting of a million characters, can be stored uncompressed in about 1 Mbyte, because each character of text occupies one byte (the Colophon at the end of the book illustrates this with data collected from the book itself).  Exercise 10.1: It is a handy rule of thumb that an average book occupies about a million bytes. Explain why this makes sense. In contrast, images occupy much more space, lending another meaning to the phrase “a picture is worth a thousand words.” Depending on the number of colors used in an image, a single pixel occupies between one bit and three bytes. Thus, a 4 Mpixel picture taken by a typical current (2006) digital camera occupies between 512 Kbytes and 12 Mbytes before compression. With the advent of powerful, inexpensive personal computers in the 1980s and 1990s came multimedia applications, where text, images, movies, and audio are stored in the computer, and can be uploaded, downloaded, displayed, edited, and played back. The storage requirements of sound are smaller than those of images or video, but greater than those of text. This is why audio compression has become important and has been the subject of much research and experimentation throughout the 1990s. Two important features of audio compression are (1) it can be lossy and (2) it requires fast decoding. Text compression must be lossless, but images and audio can lose much data without a noticeable degradation of quality. Thus, there are both lossless and lossy audio compression algorithms. It only rarely happens that a user will want to read text while it is decoded and decompressed, but this is common with audio. Often, audio is stored in compressed form and has to be decompressed in real-time when the user wants to listen to it. This is why most audio compression methods are asymmetric. The encoder can be slow, but the decoder has to be fast. This is also why audio compression methods are not dictionary based. A dictionary-based compression method may have many advantages, but fast decoding is not one of them. DOI 10.1007/978-1-84882-903-9_10, © Springer-Verlag London Limited 2010 D. Salomon, G. Motta, Handbook of Data Compression, 5th ed. 954 10. Audio Compression The field of audio compression is rapidly developing and more and more references become available. Out of this huge crowd, reference [wiki.audio 06] should especially be noted. It should be kept up-to-date and provide the reader with the latest on techniques, resources, and links to works in this important field. An important survey paper on lossless audio compression is [Hans and Schafer 01]. This chapter starts with a short introduction to sound and digitized audio. It then discusses those properties of the human auditory system (ear and brain) that make it possible to delete much audio information without adversely affecting the quality of the reconstructed audio. The chapter continues with a survey of several lossy and lossless audio compression methods. Most methods are general and can compress any audio data. Other methods are aimed at compressing speech. Of special note is the detailed description of the three lossy methods (layers) of audio compression used by the MPEG-1 and MPEG-2 standards (Section 10.14). 10.1 Sound To most of us, sound is a very familiar phenomenon, because we hear it all the time. Nevertheless, when we try to define sound, we find that we can approach this concept from two different points of view, and we end up with two definitions, as follows: An intuitive definition: Sound is the sensation detected by our ears and interpreted by our brain in a certain way. A scientific definition: Sound is a physical disturbance in a medium. It propagates in the medium as a pressure wave by the movement of atoms or molecules. We normally hear sound as it propagates through the air and hits the diaphragm in our ears. However, sound can propagate in many different media. Marine animals produce sounds underwater and respond to similar sounds. Hitting the end of a metal bar with a hammer produces sound waves that propagate through the bar and can be detected at the other end. Good sound insulators are rare, and the best insulator is vacuum, where there are no particles to vibrate and propagate the disturbance. Sound can also be considered a wave, even though its frequency may vary all the time. It is a longitudinal wave, one where the disturbance is in the direction of the wave itself. In contrast, electromagnetic waves and ocean waves are transverse waves. Their undulations are perpendicular to the direction of the wave. Like any other wave, sound has three important attributes, its speed, amplitude, and period. The frequency of a wave is not an independent attribute; it is the number of periods that occur in one time unit (one second). The unit of frequency is the hertz (Hz). The speed of sound depends mostly on the medium it passes through, and on the temperature. In air, at sea level (one atmospheric pressure), and at 20. Celsius (68. Fahrenheit), the speed of sound is 343.8 meters per second (about 1128 feet per second). The human ear is sensitive to a wide range of sound frequencies, normally from about 20 Hz to about 22,000 Hz, depending on a person’s age and health. This is the range of audible frequencies. Some animals, most notably dogs and bats, can hear higher frequencies (ultrasound). A quick calculation reveals the wavelengths associated with audible frequencies. At 22,000 Hz, each wavelength is about 1.56 cm long, whereas at 20 Hz, a wavelength is about 17.19 meters long. 10.1 Sound 955  Exercise 10.2: Verify these calculations. The amplitude of sound is also an important property. We perceive it as loudness. We sense sound when air molecules strike the diaphragm of the ear and apply pressure to it. The molecules move back and forth tiny distances that are related to the amplitude, not to the period of the sound. The period of a sound wave may be several meters, yet an individual molecule in the air may move just a millionth of a centimeter in its oscillations. With very loud noise, an individual molecule may move about one thousandth of a centimeter. A device to measure noise levels should therefore be based on a sensitive diaphragm where the pressure of the sound wave is sensed and is converted to electrical voltage, which in turn is displayed as a numeric value. The problem with measuring noise intensity is that the ear is sensitive to a very wide range of sound levels (amplitudes). The ratio between the sound level of a cannon at the muzzle and the lowest level we can hear (the threshold of hearing) is about 11–12 orders of magnitude. If we denote the lowest audible sound level by 1, then the cannon noise would have a magnitude of 1011! It is inconvenient to deal with measurements in such a wide range, which is why the units of sound loudness use a logarithmic scale. The (base-10) logarithm of 1 is 0, and the logarithm of 1011 is 11. Using logarithms, we only have to deal with numbers in the range 0 through 11. In fact, this range is too small, and we typically multiply it by 10 or by 20, to get numbers between 0 and 110 or 220. This is the well-known (and sometimes confusing) decibel system of measurement.  Exercise 10.3: (For the mathematically weak.) What exactly is a logarithm? The decibel (dB) unit is defined as the base-10 logarithm of the ratio of two physical quantities whose units are powers (energy per time). The logarithm is then multiplied by the convenient scale factor 10. (If the scale factor is not used, the result is measured in units called “Bel.” The Bel, however, was dropped long ago in favor of the decibel.) Thus, we have Level = 10 log10 P1 P2 dB, where P1 and P2 are measured in units of power such as watt, joule/sec, gram·cm/sec, or horsepower. This can be mechanical power, electrical power, or anything else. In measuring the loudness of sound, we have to use units of acoustical power. Since even loud sound can be produced with very little energy, we use the microwatt (10-6 watt) as a convenient unit. From the Dictionary Acoustics: (1) The science of sound, including the generation, transmission, and effects of sound waves, both audible and inaudible. (2) The physical qualities of a room or other enclosure (such as size, shape, amount of noise) that determine the audibility and perception of speech and music within the room. The decibel is the logarithm of a ratio. The numerator, P1, is the power (in microwatts) of the sound whose intensity level is being measured. It is convenient to select as the denominator the number of microwatts that produce the faintest audible sound (the threshold of hearing). This number is shown by experiment to be 10-6 microwatt = 10-12 watt. Thus, a stereo unit that produces 1 watt of acoustical 956 10. Audio Compression power has an intensity level of 10 log 106 10-6 = 10 log 1012 = 10Χ12 = 120 dB (this happens to be about the threshold of feeling; see Figure 10.1), whereas an earphone producing 3Χ10-4 microwatt has a level of 10 log 3Χ10-4 10-6 = 10 log 3Χ102 = 10Χ(log 3 + log 100) = 10Χ(0.477 + 2) . 24.77 dB. In the field of electricity, there is a simple relation between (electrical) power P and pressure (voltage) V . Electrical power is the product of the current and voltage P = I · V . The current, however, is proportional to the voltage by means of Ohm’s law I = V/R (where R is the resistance). We can therefore write P = V 2/R, and use pressure (voltage) in our electric decibel measurements. In practical acoustics work, we don’t always have access to the source of the sound, so we cannot measure its electrical power output. In practice, we may find ourselves in a busy location, holding a sound decibel meter in our hands, trying to measure the noise level around us. The decibel meter measures the pressure Pr applied by the sound waves on its diaphragm. Fortunately, the acoustical power per area (denoted by P) is proportional to the square of the sound pressure Pr. This is because sound power P is the product of the pressure Pr and the speed v of the sound, and because the speed can be expressed as the pressure divided by the specific impedance of the medium through which the sound propagates. This is why sound loudness is commonly measured in units of dB SPL (sound pressure level) instead of sound power. The definition is Level = 10 log10 P1 P2 = 10 log10 Pr2 1 Pr2 2 = 20 log10 Pr1 Pr2 dB SPL. The zero reference level for dB SPL becomes 0.0002 dyne/cm2, where the dyne, a small unit of force, is about 0.0010197 grams. Since a dyne equals 10-5 newtons, and since one centimeter is 0.01 meter, that zero reference level (the threshold of hearing) equals 0.00002 newton/meter2. Table 10.2 shows typical dB values in both units of power and SPL. The sensitivity of the human ear to sound level depends on the frequency. Experiments indicate that people are more sensitive to (and therefore more annoyed by) high-frequency sounds (which is why sirens have a high pitch). It is possible to modify the dB SPL system to make it more sensitive to high frequencies and less sensitive to low frequencies. This is called the dBA standard (ANSI standard S1.4-1983). There are also dBB and dBC standards of noise measurement. (Electrical engineers use also decibel standards called dBm, dBm0, and dBrn; see, for example, [Shenoi 95].) Because of the use of logarithms, dB measures don’t simply add up. If the first trumpeter starts playing his trumpet just before the concert, generating, say, a 70 dB noise level, and the second trombonist follows on his trombone, generating the same sound level, then the (poor) listeners hear twice the noise intensity, but this corresponds 10.1 Sound 957 cannon at muzzle jet aircraft threshold of feeling niagara falls factory noise office quiet home rocket threshold of pain thunder train 210 190 170 150 130 110 90 70 50 30 10 220 200 180 160 140 120 100 80 60 40 20 0 dB SPL = 0.0002 dyne/cm2 busy street full classroom recording studio isolated countryside threshold of hearing Figure 10.1: Common Sound Levels in dB SPL Units. watts dB pressure n/m2 dB SPL source 30000.0 165 2000.0 160 jet 300.0 145 200.0 140 threshold of pain 3.0 125 20.0 120 factory noise 0.03 105 2.0 100 highway traffic 0.0003 85 0.2 80 appliance 0.000003 65 0.02 60 conversation 0.00000003 45 0.002 40 quiet room 0.0000000003 25 0.0002 20 whisper 0.000000000001 0 0.00002 0 threshold of hearing Table 10.2: Sound Levels in Power and Pressure Units. to just 73 dB, not 140 dB. To show why this is true we notice that if 10 log P1 P2 = 70, then 10 log 2P1 P2 = 10log10 2 + log P1 P2  = 10(0.3 + 70/10) = 73. Doubling the noise level increases the dB level by 3 (if SPL units are used, the 3 should be doubled to 6).  Exercise 10.4: Two sound sources A and B produce dB levels of 70 and 79 dB, respectively. How much louder is source B compared to A? 958 10. Audio Compression 10.2 Digital Audio Much as an image can be digitized and broken up into pixels, where each pixel is a number, sound can also be digitized and broken up into numbers. When sound is played into a microphone, it is converted into a voltage that varies continuously with time. Figure 10.3 shows a typical example of sound that starts at zero and oscillates several times. Such voltage is the analog representation of the sound. Digitizing sound is done by measuring the voltage at many points in time, translating each measurement into a number, and writing the numbers on a file. This process is called sampling. The sound wave is sampled, and the samples become the digitized sound. The device used for sampling is called an analog-to-digital converter (ADC). The difference between a sound wave and its samples can be compared to the difference between an analog clock, where the hands seem to move continuously, and a digital clock, where the display changes abruptly every second. Since the audio samples are numbers, they are easy to edit. However, the main use of an audio file is to play it back. This is done by converting the numeric samples back into voltages that are continuously fed into a speaker. The device that does that is called a digital-to-analog converter (DAC). Intuitively, it is clear that a high sampling rate would result in better sound reproduction, but also in many more samples and therefore bigger files. Thus, the main problem in audio sampling is how often to sample a given sound. Figure 10.3a shows what may happen if the sampling rate is too low. The sound wave in the figure is sampled four times, and all four samples happen to be identical. When these samples are used to play back the sound, the result is silence. Figure 10.3b shows seven samples, and they seem to “follow” the original wave fairly closely. Unfortunately, when they are used to reproduce the sound, they produce the curve shown in dashed. There simply are not enough samples to reconstruct the original sound wave. The solution to the sampling problem is to sample sound at a little over the Nyquist frequency (page 741), which is twice the maximum frequency contained in the sound. Thus, if a sound contains frequencies of up to 2 kHz, it should be sampled at a little more than 4 kHz. (A detail that’s often ignored in the literature is that signal reconstruction is perfect only if something called the Nyquist-Shannon sampling theorem [Wikipedia 03] is used.) Such a sampling rate guarantees true reproduction of the sound. This is illustrated in Figure 10.3c, which shows 10 equally-spaced samples taken over four periods. Notice that the samples do not have to be taken from the maxima or minima of the wave; they can come from any point. Section 7.1 discusses the application of sampling to pixels in images (two-dimensional signals). The range of human hearing is typically from 16–20 Hz to 20,000–22,000 Hz, depending on the person and on age. When sound is digitized at high fidelity, it should therefore be sampled at a little over the Nyquist rate of 2Χ22000 = 44000 Hz. This is why high-quality digital sound is based on a 44,100-Hz sampling rate. Anything lower than this rate results in distortions, while higher sampling rates do not produce any improvement in the reconstruction (playback) of the sound. We can consider the sampling rate of 44,100 Hz a lowpass filter, since it effectively removes all the frequencies above 22,000 Hz. 10.2 Digital Audio 959 -1 1 1 Time Amplitude 2 3 4 0 0 -1 1 1 Amplitude 2 3 4 0 0 (a) (b) -1 1 1 Amplitude 2 3 4 0 0 (c) Figure 10.3: Sampling a Sound Wave. Many low-fidelity applications sample sound at 11,000 Hz, and the telephone system, originally designed for conversations, not for digital communications, samples sound at only 8 kHz. Thus, any frequency higher than 4000 Hz gets distorted when sent over the phone, which is why it is hard to distinguish, on the phone, between the sounds of f and s. This is also why, when someone gives you an address over the phone you should ask, “Is it H street, as in EFGH?” Often, the answer is, “No, this is Eighth street, as in sixth, seventh, eighth.” The meeting was in Mr. Rogers’ law office, at 1415 H Street. My slip of paper said 1415 8th Street. (The address had been given over the telephone.) —Richard P. Feynman, What Do YOU Care What Other People Think? The second problem in sound sampling is the sample size. Each sample becomes a number, but how large should this number be? In practice, samples are normally either 8 960 10. Audio Compression or 16 bits, although some high-quality sound cards that are available for many computer platforms may optionally use 32-bit samples. Assuming that the highest voltage in a sound wave is 1 volt, an 8-bit sample can distinguish voltages as low as 1/256 . 0.004 volt, or 4 millivolts (mv). A quiet sound, generating a wave lower than 4 mv, would be sampled as zero and played back as silence. In contrast, with a 16-bit sample it is possible to distinguish sounds as low as 1/65, 536 . 15 microvolt (΅v). We can think of the sample size as a quantization of the original audio data. Eight-bit samples are more coarsely quantized than 16-bit samples. As a result, they produce better compression but poorer audio reconstruction (the reconstructed sound features only 256 different amplitudes).  Exercise 10.5: Suppose that the sample size is one bit. Each sample has a value of either 0 or 1. What would we hear when these samples are played back? Audio sampling is also called pulse code modulation (PCM).We have all heard of AM and FM radio. These terms stand for amplitude modulation and frequency modulation, respectively. They indicate methods to modulate (i.e., to include binary information in) continuous waves. The term pulse modulation refers to techniques for converting a continuous wave to a stream of binary numbers (audio samples). Possible pulse modulation methods include pulse amplitude modulation (PAM), pulse position modulation (PPM), pulse width modulation (PWM), and pulse number modulation (PNM). [Pohlmann 85] is a good source of information on these methods. In practice, however, PCM has proved the most effective form of converting sound waves to numbers. When stereo sound is digitized, the PCM encoder multiplexes the left and right sound samples. Thus, stereo sound sampled at 22,000 Hz with 16-bit samples generates 44,000 16-bit samples per second, for a total of 704,000 bits/sec, or 88,000 bytes/sec. 10.2.1 Digital Audio and Laplace Distribution The Laplace distribution has already been mentioned in Section 7.25. This short section explains the relevance of this distribution to audio compression. A large audio file with a long, complex piece of music tends to have all the possible values of audio samples. Consider the simple case of 8-bit audio samples, which have values in the interval [0, 255]. A large audio file, with millions of audio samples, will tend to have many audio samples concentrated around the center of this interval (around 128), fewer large samples (close to the maximum 255), and few small samples (although there may be many audio samples of 0, because many types of sound tend to have periods of silence). The distribution of the samples may have a maximum at its center and another spike at 0. Thus, the audio samples themselves do not normally have a simple distribution. However, when we examine the differences of adjacent samples, we observe a completely different behavior. Consecutive audio samples tend to be correlated, which is why the differences of consecutive samples tend to be small numbers. Experiments with many types of sound indicate that the distribution of audio differences resembles the Laplace distribution. The following experiment reinforces this conclusion. Figure 10.4a shows the distribution of about 2900 audio samples taken from a real audio file (file a8.wav, where the single English letter a is spoken). It is easy to see that the distribution has a peak at 128, but is very rough and does not resemble either the Gaussian or the Laplacian distributions. Also, there are very few samples smaller than 75 and 10.3 The Human Auditory System 961 greater than 225 (because of the small size of the file and its particular, single-letter, content). However, part b of the figure is different. It displays the distribution of the differences and it is clearly smooth, narrow, and has a sharp peak at 0 (in the figure, the peak is at 256 because indexes in Matlab must be nonnegative, which is why an artificial 256 is added to the indexes of differences). Parts c,d of the figure show the distribution of about 2900 random numbers. The numbers themselves have a flat (although nonsmooth) distribution, while their differences have a very rough distribution, centered at 0 but very different from Laplace (only about 24 differences are zero). The distribution of the differences is not flat; it resembles a very rough Gaussian distribution. Given random integers Ri between 0 and n, the differences of consecutive numbers can be from -n to +n, but these differences occur with different probabilities. Extreme differences are rare. The only way to obtain a difference of n is to have an Ri+1 = 0 follow an Ri = n and subtract n - 0. Small differences, on the other hand, have higher probabilities because they can be obtained in more cases. A difference of 2 is obtained by subtracting 56 - 54, but also 57 - 55, 58 - 56 and so on. The conclusion is that differences of consecutive correlated values tend to have a narrow, peaked distribution, resembling the Laplace distribution. This is true for the differences of audio samples as well as for the differences of consecutive pixels of an image. A compression algorithm may take advantage of this fact and encode the differences with variable-length codes that have a Laplace distribution. A more sophisticated version may compute differences between actual values (audio samples or pixels) and their predicted values, and then encode the (Laplace distributed) differences. Two such methods are image MLP (Section 7.25) and FLAC (Section 10.10). 10.3 The Human Auditory System The frequency range of the human ear is from about 20 Hz to about 20,000 Hz, but the ear’s sensitivity to sound is not uniform. It depends on the frequency, and experiments indicate that in a quiet environment the ear’s sensitivity is maximal for frequencies in the range 2 kHz to 4 kHz. Figure 10.5a shows the hearing threshold for a quiet environment (see also Figure 10.67).  Exercise 10.6: Come up with an appropriate way to conduct such experiments. It should also be noted that the range of the human voice is much more limited. It is only from about 500 Hz to about 2 kHz (Section 10.8). The existence of the hearing threshold suggests an approach to lossy audio compression. Just delete any audio samples that are below the threshold. Since the threshold depends on the frequency, the encoder needs to know the frequency spectrum of the sound being compressed at any time. The encoder therefore has to save several of the previously input audio samples at any time (n-1 samples, where n is either a constant or a user-controlled parameter). When the current sample is input, the first step is to transform the most-recent n samples to the frequency domain (Section 8.2). The result is a number m of values (called signals) that indicate the strength of the sound at m 962 10. Audio Compression 0 50 100 150 200 250 300 0 10 20 30 40 50 60 70 80 0 100 200 300 400 500 600 0 20 40 60 80 100 120 140 160 0 50 100 150 200 250 300 4 6 8 10 12 14 16 18 20 0 100 200 300 400 500 600 0 5 10 15 20 25 (a) (b) (c) (d) Figure 10.4: Distribution of Audio Samples, Random Numbers, and Differences. % File ’LaplaceWav.m’ filename=’a8.wav’; dim=2950; % size of file a8.wav dist=zeros(256,1); ddist=zeros(512,1); fid=fopen(filename,’r’); buf=fread(fid,dim,’uint8’); %input unsigned integers for i=46:dim % skip .wav file header x=buf(i)+1; dif=buf(i)-buf(i-1)+256; dist(x)=dist(x)+1; ddist(dif)=ddist(dif)+1; end figure(1), plot(dist), colormap(gray) %dist of audio samples figure(2), plot(ddist), colormap(gray) %dist of differences dist=zeros(256,1); ddist=zeros(512,1); % clear buffers buf=randint(dim,1,[0 255]); % many random numbers for i=2:dim x=buf(i)+1; dif=buf(i)-buf(i-1)+256; dist(x)=dist(x)+1; ddist(dif)=ddist(dif)+1; end figure(3), plot(dist), colormap(gray) %dist of random numbers figure(4), plot(ddist), colormap(gray) %dist of differences Code for Figure 10.4. 10.3 The Human Auditory System 963 0 2 20 40 60 80 4 6 8 10 12 14 14 18 20 22 23 24 25 16 (c) dB KHz Bark 0 2 10 20 30 40 4 6 8 10 12 14 16 (a) dB KHz Frequency 0 2 10 20 30 40 4 6 8 10 12 14 16 (b) dB KHz Frequency x Figure 10.5: Threshold and Masking of Sound. 964 10. Audio Compression different frequencies. If a signal for frequency f is smaller than the hearing threshold at f, it (the signal) should be deleted. In addition to this, two more properties of the human hearing system are used in audio compression. They are frequency masking and temporal masking. Frequency masking (also known as auditory masking) occurs when a sound that we can normally hear (because it is loud enough) is masked by another sound with a nearby frequency. The thick arrow in Figure 10.5b represents a strong sound source at 800 kHz. This source raises the normal threshold in its vicinity (the dashed curve), with the result that the nearby sound represented by the arrow at “x”, a sound that would normally be audible because it is above the threshold, is now masked, and is inaudible (see also Figure 10.67). A good lossy audio compression method should identify this case and delete the signals corresponding to sound “x”, because it cannot be heard anyway. This is one way to lossily compress sound. The frequency masking (the width of the dashed curve of Figure 10.5b) depends on the frequency. It varies from about 100 Hz for the lowest audible frequencies to more than 4 kHz for the highest. The range of audible frequencies can therefore be partitioned into a number of critical bands that indicate the declining sensitivity of the ear (rather, its declining resolving power) for higher frequencies. We can think of the critical bands as a measure similar to frequency. However, in contrast to frequency, which is absolute and has nothing to do with human hearing, the critical bands are determined according to the sound perception of the ear. Thus, they constitute a perceptually uniform measure of frequency. Table 10.6 lists 27 approximate critical bands. Another way to describe critical bands is to say that because of the ear’s limited perception of frequencies, the threshold at a frequency f is raised by a nearby sound only if the sound is within the critical band of f. This also points the way to designing a practical lossy compression algorithm. The audio signal should first be transformed into its frequency domain, and the resulting values (the frequency spectrum) should be divided into subbands that resemble the critical bands as much as possible. Once this is done, the signals in each subband should be quantized such that the quantization noise (the difference between the original sound sample and its quantized value) should be inaudible. band range band range band range 0 0–50 9 800–940 18 3280–3840 1 50–95 10 940–1125 19 3840–4690 2 95–140 11 1125–1265 20 4690–5440 3 140–235 12 1265–1500 21 5440–6375 4 235–330 13 1500–1735 22 6375–7690 5 330–420 14 1735–1970 23 7690–9375 6 420–560 15 1970–2340 24 9375–11625 7 560–660 16 2340–2720 25 11625–15375 8 660–800 17 2720–3280 26 15375–20250 Table 10.6: Twenty-Seven Approximate Critical Bands. 10.3 The Human Auditory System 965 Yet another way to look at the concept of critical bands is to consider the human auditory system a filter that lets through only frequencies in the range (bandpass) of 20 Hz to 20000 Hz. We visualize the ear–brain system as a collection of filters, each with a different bandpass. The bandpasses are called critical bands. They overlap and they have different widths. They are narrow (about 100 Hz) at low frequencies and become wider (to about 4–5 kHz) at high frequencies. The width of a critical band is called its size. The widths of the critical bands introduce a new unit, the Bark (after H. G. Barkhausen) such that one Bark is the width (in Hz) of one critical band. The Bark is defined as 1 Bark = " f 100 , for frequencies f < 500 Hz, 9 + 4 log2  f 1000, for frequencies f . 500 Hz. Figure 10.5c shows some critical bands, with Barks between 14 and 25, positioned above the threshold. Heinrich Georg Barkhausen Heinrich Barkhausen was born on December 2, 1881, in Bremen, Germany. He spent his entire career as a professor of electrical engineering at the Technische Hochschule in Dresden, where he concentrated on developing electron tubes. He also discovered the so-called “Barkhausen effect,” where acoustical waves are generated in a solid by the movement of domain walls when the material is magnetized. He also coined the term “phon” as a unit of sound loudness. The institute in Dresden was destroyed, as was most of the city, in the famous fire bombing in February 1945. After the war, Barkhausen helped rebuild the institute. He died on February 20, 1956. 966 10. Audio Compression Temporal masking may occur when a strong sound A of frequency f is preceded or followed in time by a weaker sound B at a nearby (or the same) frequency. If the time interval between the sounds is short, sound B may not be audible. Figure 10.7 illustrates an example of temporal masking (see also Figure 10.68). The threshold of temporal masking due to a loud sound at time 0 goes down, first sharply, then slowly. A weaker sound of 30 dB will not be audible if it occurs 10 ms before or after the loud sound, but will be audible if the time interval between the sounds is 20 ms. 20 30 40 60 -500 -200-100 -50 -20 -10 0 10 20 50 100 200 500 dB ms Time Figure 10.7: Threshold and Masking of Sound. 10.3.1 Conventional Methods Conventional compression methods, such as RLE, statistical, and dictionary-based, can be used to losslessly compress audio data, but the results depend heavily on the specific sound. Some sounds may compress well under RLE but not under a statistical method. Other sounds may lend themselves to statistical compression but may expand when processed by a dictionary method. Here is how sounds respond to each of the three classes of compression methods. RLE may work well when the sound contains long runs of identical samples. With 8-bit samples this may be common. Recall that the difference between the two 8-bit samples n and n + 1 is about 4 mv. A few seconds of uniform music, where the wave does not oscillate more than 4 mv, may produce a run of thousands of identical samples. With 16-bit samples, long runs may be rare and RLE, consequently, ineffective. Statistical methods assign variable-length codes to the samples according to their frequency of occurrence. With 8-bit samples, there are only 256 different samples, so in a large audio file, the samples may sometimes have a flat distribution. Such a file will therefore not respond well to Huffman coding (See Exercise 5.7). With 16-bit samples there are more than 65,000 sample values, so they may sometimes feature skewed probabilities (i.e., some samples may occur very often, while others may be rare). Such a file may therefore compress better with arithmetic coding, which works well even for skewed probabilities. Dictionary-based methods expect to find the same phrases again and again in the data. This happens with text, where certain strings may repeat often. Sound, however, is an analog signal and the particular samples generated depend on the precise way the 10.3 The Human Auditory System 967 ADC works. With 8-bit samples, for example, a wave of 8 mv becomes a sample of size 2, but waves very close to that, say, 7.6 mv or 8.5 mv, may become samples of different sizes. This is why parts of speech that sound the same to us, and should therefore have become identical phrases, end up being digitized slightly differently, and go into the dictionary as different phrases, thereby reducing compression. Dictionary-based methods are not well suited for sound compression. I don’t like the sound of that sound. —Heather Graham as Judy Robinson in Lost in Space (1998) 10.3.2 Lossy Audio Compression It is possible to get better audio compression by developing lossy methods that take advantage of our perception of sound, and discard data to which the human ear is not sensitive. This is similar to lossy image compression, where visual data to which the human eye is not sensitive is discarded. In both cases we use the fact that the original information (image or sound) is analog and has already lost some quality when digitized. Losing some more data, if done carefully, may not significantly affect the played-back sound, and may therefore be indistinguishable from the original. We briefly describe two approaches, silence compression and companding. The principle of silence compression is to treat small samples as if they were silence (i.e., as samples of 0). This generates run lengths of zero, so silence compression is actually a variant of RLE, suitable for sound compression. This method uses the fact that some people have less sensitive hearing than others, and will tolerate the loss of sound that is so quiet they may not hear it anyway. Audio files containing long periods of low-volume sound will respond to silence compression better than other files with high-volume sound. This method requires a user-controlled parameter that specifies the largest sample that should be suppressed. Two other parameters are also necessary, although they may not have to be user-controlled. One specifies the shortest run length of small samples, typically 2 or 3. The other specifies the minimum number of consecutive large samples that should terminate a run of silence. For example, a run of 15 small samples, followed by two large samples, followed by 13 small samples may be considered one silence run of 30 samples, whereas the runs 15, 3, 13 may become two distinct silence runs of 15 and 13 samples, with nonsilence in between. Companding (short for “compressing/expanding”) uses the fact that the ear requires more precise samples at low amplitudes (soft sounds), but is more forgiving at higher amplitudes. A typical ADC used in sound cards for personal computers converts voltages to numbers linearly. If an amplitude a is converted to the number n, then amplitude 2a will be converted to the number 2n. A compression method using companding examines every sample in the sound file, and employs a nonlinear formula to reduce the number of bits devoted to it. For 16-bit samples, for example, a companding encoder may use a formula as simple as mapped = 32,767 2sample 65536 - 1 (10.1) to reduce each sample. This formula maps the 16-bit samples nonlinearly to 15-bit numbers (i.e., numbers in the range [0, 32767]) such that small samples are less affected 968 10. Audio Compression than large ones. Table 10.8 illustrates the nonlinearity of this mapping. It shows eight pairs of samples, where the two samples in each pair differ by 100. The two samples of the first pair get mapped to numbers that differ by 34, whereas the two samples of the last pair are mapped to numbers that differ by 65. The mapped 15-bit numbers can be decoded back into the original 16-bit samples by the inverse formula Sample = 65,536 log2 1 + mapped 32,767 . (10.2) Sample Mapped Diff Sample Mapped Diff 100 › 35 30,000 › 12,236 200 › 69 34 30,100 › 12,283 47 1,000 › 348 40,000 › 17,256 1,100 › 383 35 40,100 › 17,309 53 10,000 › 3,656 50,000 › 22,837 10,100 › 3,694 38 50,100 › 22,896 59 20,000 › 7,719 60,000 › 29,040 20,100 › 7,762 43 60,100 › 29,105 65 Table 10.8: 16-Bit Samples Mapped to 15-Bit Numbers. Reducing 16-bit numbers to 15 bits doesn’t produce much compression. Better compression can be achieved by substituting a smaller number for 32,767 in equations (10.1) and (10.2). A value of 127, for example, would map each 16-bit sample into an 8-bit one, yielding a compression ratio of 0.5. However, decoding would be less accurate. A 16-bit sample of 60,100, for example, would be mapped into the 8-bit number 113, but this number would produce 60,172 when decoded by Equation (10.2). Even worse, the small 16-bit sample 1000 would be mapped into 1.35, which has to be rounded to 1. When Equation (10.2) is used to decode a 1, it produces 742, significantly different from the original sample. The amount of compression should therefore be a user-controlled parameter, and this is an interesting example of a compression method where the compression ratio is known in advance! In practice, there is no need to go through Equations (10.1) and (10.2), since the mapping of all the samples can be prepared in advance in a table. Both encoding and decoding are therefore fast. Companding is not limited to Equations (10.1) and (10.2). More sophisticated methods, such as ΅-law and A-law, are commonly used and have been designated international standards. 10.4 WAVE Audio Format 969 10.4WAVE Audio Format WAVE (or simply Wave) is the native file format employed by the Windows operating system for storing digital audio data. Audio has become as important as images and video, which is why modern operating systems support a native format for audio files. The popularity of Windows and the vast amount of software available for the PC platform have guaranteed the popularity of the Wave format, hence this short description. First, three technical notes. (1) The Wave file format is native to Windows and therefore runs on Intel processors which use the little endian byte order. (2) Wave files normally contain strings of text for specifying cue points, labels, notes, and other information. Strings whose size is not known in advance are stored in the so-called Pascal format where the first byte specifies the number of the ASCII text bytes in the string. (3) The WAV (or .WAV) file format is a special case of Wave where no compression is used. A recommended reference for Wave and other important file formats is [Born 95]. However, it is always a good idea to search the Internet for new references. The organization of a Wave file is based on the standard RIFF structure. (RIFF stands for Resource Interchange File Format. This is the basis of several important file formats.) The file consists of chunks, where each chunk starts with a header and may contain subchunks as well as data. For example, the fmt and data chunks are actually subchunks of the RIFF chunk. New types of chunks may be added in the future, leading to a situation where existing software does not recognize new chunks. Thus, the general rule is to skip any unrecognizable chunks. The total size of a chunk must be an even number of bytes (a multiple of two bytes), which is why a chunk may end with one byte of zero padding. The first chunk of a Wave file is RIFF. It start with the 4-byte string RIFF, followed by the size of the remaining part of the file (in four bytes), followed by string WAVE. This is immediately followed by the format and data subchunks. The former is listed in Table 10.9. The main specifications included in the format chunk are as follows: Compression Code: Wave files may be compressed. A special case is .WAV files, which are not compressed. The compression codes are listed in Table 10.10. Number of Channels: The number of audio channels encoded in the data chunk. A value of 1 indicates a mono signal, 2 indicates stereo, and so on. Sample Rate: The number of audio samples per second. This value is unaffected by the number of channels. Average Bytes Per Second: How many bytes of wave data must be sent to a D/A converter per second in order to play the wave file? This value is the product of the sample rate and block align. Block Align: The number of bytes per sample slice. This value is not affected by the number of channels. Significant Bits Per Sample: The number of bits per audio sample, most often 8, 16, 24, or 32. If the number of bits is not a multiple of 8, then the number of bytes used 970 10. Audio Compression Byte # Description 0–3 fmt 4–7 Length of subchunk (24+extra format bytes) 8–9 16-bit compression code 10–11 Channel numbers (1=Mono, 2=Stereo) 12-15 Sample Rate (Binary, in Hz) 16–19 Bytes Per Second 20–21 Bytes Per Sample: 1=8 bit Mono, 2=8 bit Stereo or 16 bit Mono, 4=16 bit Stereo 22-23 Bits Per Sample 24– Extra format bytes Table 10.9: A Format Chunk. Code (Hex) Compression 0 0000 Unknown 1 0001 PCM/uncompressed 2 0002 Microsoft ADPCM 6 0006 ITU G.711 A-law 7 0007 ITU G.711 ΅-law 17 0011 IMA ADPCM 20 0016 ITU G.723 ADPCM (Yamaha) 49 0031 GSM 6.10 64 0040 ITU G.721 ADPCM 80 0050 MPEG 65,536 FFFF Experimental Table 10.10: Wave Compression Codes. per sample is rounded up to the nearest byte size and the unused bytes are set to 0 and ignored. Extra Format Bytes: The number of additional format bytes that follow. These bytes are used for nonzero compression codes. Their meaning depends on the specific compression used. If this value is not even (a multiple of 2), a single byte of zero-padding should be added to the end of this data to align it, but the value itself should remain odd. The data chunk starts with the string data, which is followed by the size of the data that follows (as a 32-bit integer). This is followed by the data bytes (audio samples). Table 10.11 lists the first 80 bytes of a simple .WAV file. The file starts with the identifying string RIFF. The next four bytes are the length, which is 16e616 = 5,86210 bytes. This number is the length of the entire file minus the eight bytes for the RIFF and length. The strings WAVE and fmt follow. The second row starts with the length 1016 = 2410 of the format chunk. This indicates no extra bytes. The next two bytes are 1 (compression code of uncompressed), followed by another two bytes of 1 (mono audio, only one channel). The audio sampling 10.5 ΅-Law and A-Law Companding 971 Hexadecimal ASCII 00 52 49 46 46 E6 16 00 00 57 41 56 45 66 6D 74 20 RIFF....WAVEfmt 10 10 00 00 00 01 00 01 00 10 27 00 00 20 4E 00 00 .........’...N.. 20 02 00 10 00 64 61 74 61 C2 16 00 00 8B FC 12 F8 ....data........ 30 B9 F8 EB F8 25 F8 31 F3 67 EE 66 ED 80 EB F1 E7 ....%.1.g.f..... 40 3D EB 5C 09 C3 1B 44 E9 EF FF 18 39 35 0E 5B 0A =.\...D....95.[. Table 10.11: Example of a .WAV File. rate is 2,71016 = 10,00010, and the bytes per second are 4e2016 = 20,00010. The third row starts with bytes per sample (two bytes of 2, indicating 8-bit stereo or 16-bit mono). The next two bytes are the bits per sample (16 in our case). We now know that this audio is mono with 16-bit audio samples. The data subchunk follows immediately. It starts with the string data, followed by the four bytes 16c216 indicating that 582610 bytes of audio data follow. The first four such bytes (two audio samples) are 8B FC 12 F8. 10.5 ΅-Law and A-Law Companding These two international standards, formally known as recommendation G.711, are documented in [ITU-T 89]. They employ logarithm-based functions to encode audio samples for ISDN (integrated services digital network) digital telephony services, by means of nonlinear quantization. The ISDN hardware samples the voice signal from the telephone 8,000 times per second, and generates 14-bit samples (13 for A-law). The method of ΅-law companding is used in North America and Japan, and A-law is used elsewhere. The two methods are similar; they differ mostly in their quantizations (midtread vs. midriser). Experiments (documented in [Shenoi 95]) indicate that the low amplitudes of speech signals contain more information than the high amplitudes. This is why nonlinear quantization makes sense. Imagine an audio signal sent on a telephone line and digitized to 14-bit samples. The louder the conversation, the higher the amplitude, and the bigger the value of the sample. Since high amplitudes are less important, they can be coarsely quantized. If the largest sample, which is 214 - 1 = 16,383, is quantized to 255 (the largest 8-bit number), then the compression factor is 14/8 = 1.75. When decoded, a code of 255 will become very different from the original 16,383. We say that because of the coarse quantization, large samples end up with high quantization noise. Smaller samples should be finely quantized, so they end up with low quantization noise. The ΅-law encoder inputs 14-bit samples and outputs 8-bit codewords. The Alaw inputs 13-bit samples and also outputs 8-bit codewords. The telephone signals are sampled at 8 kHz (8,000 times per second), so the ΅-law encoder receives 8,000Χ14 = 112,000 bits/sec. At a compression factor of 1.75, the encoder outputs 64,000 bits/sec. The G.711 standard [G.711 72] also specifies output rates of 48 Kbps and 56 Kbps. The ΅-law encoder receives a 14-bit signed input sample x. Thus, the input is in the range [-8, 192, +8, 191]. The sample is normalized to the interval [-1, +1], and the 972 10. Audio Compression encoder uses the logarithmic expression sgn(x) ln(1 + ΅|x|) ln(1 + ΅) , where sgn(x) = "+1, x>0, 0, x= 0, -1, x<0 (and ΅ is a positive integer), to compute and output an 8-bit code in the same interval [-1, +1]. The output is then scaled to the range [-256, +255]. Figure 10.12 shows this output as a function of the input for the three ΅ values 25, 255, and 2555. It is clear that large values of ΅ cause coarser quantization for larger amplitudes. Such values allocate more bits to the smaller, more important, amplitudes. The G.711 standard recommends the use of ΅ = 255. The diagram shows only the nonnegative values of the input (i.e., from 0 to 8191). The negative side of the diagram has the same shape but with negative inputs and outputs. The A-law encoder uses the similar expression .. .. . sgn(x) A|x| 1 + ln(A) , for 0 . |x| < 1 A, sgn(x)1 + ln(A|x|) 1 + ln(A) , for 1 A . |x| < 1. The G.711 standard recommends the use of A = 87.6. The following simple examples illustrate the nonlinear nature of the ΅-law. The two (normalized) input samples 0.15 and 0.16 are transformed by ΅-law to outputs 0.6618 and 0.6732. The difference between the outputs is 0.0114. On the other hand, the two input samples 0.95 and 0.96 (bigger inputs but with the same difference) are transformed to 0.9908 and 0.9927. The difference between these two outputs is 0.0019; much smaller. Bigger samples are decoded with more noise, and smaller samples are decoded with less noise. However, the signal-to-noise ratio (SNR, Section 7.4.2) is constant because both the ΅-law and the SNR use logarithmic expressions. Logarithms are slow to compute, so the ΅-law encoder performs much simpler calculations that produce an approximation. The output specified by the G.711 standard is an 8-bit codeword whose format is shown in Figure 10.13. Bit P in Figure 10.13 is the sign bit of the output (same as the sign bit of the 14-bit signed input sample). Bits S2, S1, and S0 are the segment code, and bits Q3 through Q0 are the quantization code. The encoder determines the segment code by (1) adding a bias of 33 to the absolute value of the input sample, (2) determining the bit position of the most significant 1-bit among bits 5 through 12 of the input, and (3) subtracting 5 from that position. The 4-bit quantization code is set to the four bits following the bit position determined in step 2. The encoder ignores the remaining bits of the input sample, and it inverts (1’s complements) the codeword before it is output. We use the input sample -656 as an example. The sample is negative, so bit P becomes 1. Adding 33 to the absolute value of the input yields 689 = 00010101100012 (Figure 10.14). The most significant 1-bit in positions 5 through 12 is found at position 9. The segment code is thus 9 - 5 = 4. The quantization code is the four bits 0101 at positions 8–5, and the remaining five bits 10001 are ignored. The 8-bit codeword (which is later inverted) becomes 10.5 ΅-Law and A-Law Companding 973 0 2000 4000 6000 8000 16 48 64 32 80 112 25 255 2555 128 96 Figure 10.12: The ΅-Law for ΅ Values of 25, 255, and 2555. dat=linspace(0,1,1000); mu=255; plot(dat*8159,128*log(1+mu*dat)/log(1+mu)); Matlab code for Figure 10.12. Notice how the input is normalized to the range [0, 1] before the calculations, and how the output is scaled from the interval [0, 1] to [0, 128]. P S2 S1 S0 Q3 Q2 Q1 Q0 Figure 10.13: G.711 ΅-Law Codeword. Q3 Q2 Q1 Q0 0 0 0 1 0 1 0 1 1 0 0 0 1 12 11 10 9 8 7 6 5 4 3 2 1 0 Figure 10.14: Encoding Input Sample -656. 974 10. Audio Compression P S2 S1 S0 Q3 Q2 Q1 Q0 1 1 0 0 0 1 0 1 The ΅-law decoder inputs an 8-bit codeword and inverts it. It then decodes it as follows: 1. Multiply the quantization code by 2 and add 33 (the bias) to the result. 2. Multiply the result by 2 raised to the power of the segment code. 3. Decrement the result by the bias. 4. Use bit P to determine the sign of the result. Applying these steps to our example produces 1. The quantization code is 1012 = 5, so 5 Χ 2 + 33 = 43. 2. The segment code is 1002 = 4, so 43 Χ 24 = 688. 3. Decrement by the bias 688 - 33 = 655. 4. Bit P is 1, so the final result is -655. Thus, the quantization error (the noise) is 1; very small. Figure 10.15a illustrates the nature of the ΅-law midtread quantization. Zero is one of the valid output values, and the quantization steps are centered at the input value of 0. The steps are organized in eight segments of 16 steps each. The steps within each segment have the same width, but they double in width from one segment to the next. If we denote the segment number by i (where i = 0, 1, . . . , 7) and the width of a segment by k (where k = 1, 2, . . . , 16), then the middle of the tread of each step in Figure 10.15a (i.e., the points labeled xj) is given by x(16i + k) = T(i) + kΧD(i), (10.3) where the constants T(i) and D(i) are the initial value and the step size for segment i, respectively. They are given by i 0 1 2 3 4 5 6 7 T(i) 1 35 103 239 511 1055 2143 4319 D(i) 2 4 8 16 32 64 128 256 Table 10.16 lists some values of the breakpoints (points xj) and outputs (points yj) shown in Figure 10.15a. The operation of the A-law encoder is similar, except that the quantization (Figure 10.15b) is of the midriser variety. The breakpoints xj are given by Equation (10.3), but the initial value T(i) and the step size D(i) for segment i are different from those used by the ΅-law encoder and are given by i 0 1 2 3 4 5 6 7 T(i) 0 32 64 128 256 512 1024 2048 D(i) 2 2 4 8 16 32 64 128 Table 10.17 lists some values of the breakpoints (points xj) and outputs (points yj) shown in Figure 10.15b. 10.5 ΅-Law and A-Law Companding 975 (a) x input y output x1 y1 y0 y2 y3 y4 x2 x3 x4 (b) x input y output x1 y1 -y2 -y1 y2 y3 y4 -x2 -x2 x2 x3 x4 Figure 10.15: (a) ΅-Law Midtread Quantization. (b) A-Law Midriser Quantization. segment 0 segment 1 · · · segment 7 break output break output break output points values points values points values y0 = 0 y16 = 33 · · · y112 = 4191 x1 = 1 x17 = 35 x113 = 4319 y1 = 2 y17 = 37 · · · y113 = 4447 x2 = 3 x18 = 39 x114 = 4575 y2 = 4 y18 = 41 · · · y114 = 4703 x3 = 5 x19 = 43 x115 = 4831 y3 = 6 y19 = 45 · · · y115 = 4959 x4 = 7 x20 = 47 x116 = 5087 · · · · · · · · · · · · x15 = 29 x31 = 91 x127 = 7903 y15 = 28 y31 = 93 · · · y127 = 8031 x16 = 31 x32 = 95 x128 = 8159 Table 10.16: Specification of the ΅-Law Quantizer. The A-law encoder generates an 8-bit codeword with the same format as the ΅-law encoder. It sets the P bit to the sign of the input sample. It then determines the segment code in the following steps: 1. Determine the bit position of the most significant 1-bit among the seven most significant bits of the input. 2. If such a 1-bit is found, the segment code becomes that position minus 4. Otherwise, the segment code becomes zero. The 4-bit quantization code is set to the four bits following the bit position determined in step 1, or to half the input value if the segment code is zero. The encoder ignores the remaining bits of the input sample, and it inverts bit P and the even-numbered bits of the codeword before it is output. 976 10. Audio Compression segment 0 segment 1 · · · segment 7 break output break output break output points values points values points values x0 = 0 x16 = 32 x112 = 2048 y1 = 1 y17 = 33 · · · y113 = 2112 x1 = 2 x17 = 34 x113 = 2176 y2 = 3 y18 = 35 · · · y114 = 2240 x2 = 4 x18 = 36 x114 = 2304 y3 = 5 y19 = 37 · · · y115 = 2368 x3 = 6 x19 = 38 x115 = 2432 y4 = 7 y20 = 39 · · · y116 = 2496 · · · · · · · · · · · · x15 = 30 x31 = 62 x128 = 4096 y16 = 31 y32 = 63 · · · y127 = 4032 Table 10.17: Specification of the A-Law Quantizer. The A-law decoder decodes an 8-bit codeword into a 13-bit audio sample as follows: 1. It inverts bit P and the even-numbered bits of the codeword. 2. If the segment code is nonzero, the decoder multiplies the quantization code by 2 and increments this by the bias (33). The result is then multiplied by 2 and raised to the power of the (segment code minus 1). If the segment code is 0, the decoder outputs twice the quantization code, plus 1. 3. Bit P is then used to determine the sign of the output. Normally, the output codewords are generated by the encoder at the rate of 64 Kbps. The G.711 standard [G.711 72] also provides for two other rates, as follows: 1. To achieve an output rate of 48 Kbps, the encoder masks out the two least-significant bits of each codeword. This works, because 6/8 = 48/64. 2. To achieve an output rate of 56 Kpbs, the encoder masks out the least-significant bit of each codeword. This works, because 7/8 = 56/64 = 0.875. This applies to both the ΅-law and the A-law. The decoder typically fills up the masked bit positions with zeros before decoding a codeword. 10.6 ADPCM Audio Compression 977 10.6 ADPCM Audio Compression Compression is possible only because audio, and therefore audio samples, tend to have redundancies. Adjacent audio samples tend to be similar in much the same way that neighboring pixels in an image tend to have similar colors. The simplest way to exploit this redundancy is to subtract adjacent samples and code the differences, which tend to be small integers. Any audio compression method based on this principle is called DPCM (differential pulse code modulation). Such methods, however, are inefficient, because they do not adapt themselves to the varying magnitudes of the audio stream. Better results are achieved by an adaptive version, and any such version is called ADPCM [ITU-T 90]. ADPCM employs linear prediction (this is also commonly used in predictive image compression). It uses the previous sample (or several previous samples) to predict the current sample. It then computes the difference between the current sample and its prediction, and quantizes the difference. For each input sample X[n], the output C[n] of the encoder is simply a certain number of quantization levels. The decoder multiplies this number by the quantization step (and may add half the quantization step, for better precision) to obtain the reconstructed audio sample. The method is efficient because the quantization step is updated all the time, by both encoder and decoder, in response to the varying magnitudes of the input samples. It is also possible to modify adaptively the prediction algorithm. Various ADPCM methods differ in the way they predict the current audio sample and in the way they adapt to the input (by changing the quantization step size and/or the prediction method). In addition to the quantized values, an ADPCM encoder can provide the decoder with side information. This information increases the size of the compressed stream, but this degradation is acceptable to the users, because it makes the compressed audio data more useful. Typical applications of side information are (1) help the decoder recover from errors and (2) signal an entry point into the compressed stream. An original audio stream may be recorded in compressed form on a medium such as a CD-ROM. If the user (listener) wants to listen to song 5, the decoder can use the side information to quickly find the start of that song. Figure 10.18a,b shows the general organization of the ADPCM encoder and decoder. Notice that they share two functional units, a feature that helps in both software and hardware implementations. The adaptive quantizer receives the difference D[n] between the current input sample X[n] and the prediction Xp[n - 1]. The quantizer computes and outputs the quantized code C[n] of X[n]. The same code is sent to the adaptive dequantizer (the same dequantizer used by the decoder), which produces the next dequantized difference value Dq[n]. This value is added to the previous predictor output Xp[n - 1], and the sum Xp[n] is sent to the predictor to be used in the next step. Better prediction would be obtained by feeding the actual input X[n] to the predictor. However, the decoder wouldn’t be able to mimic that, since it does not have X[n]. We see that the basic ADPCM encoder is simple, and the decoder is even simpler. It inputs a code C[n], dequantizes it to a difference Dq[n], which is added to the preceding predictor output Xp[n - 1] to form the next output Xp[n]. The next output is also fed into the predictor, to be used in the next step. 978 10. Audio Compression X[n] D[n] C[n] C[n] Xp[n-1] Xp[n-1] Xp[n] Xp[n] Dq[n] Dq[n] (a) (b) adaptive quantizer + + + adaptive predictor adaptive predictor quantizer adaptive dequantizer adaptive + + + - Figure 10.18: (a) ADPCM Encoder and (b) Decoder. The remainder of this section describes the particular ADPCM algorithm adopted by the Interactive Multimedia Association [IMA 06]. The IMA is a consortium of computer hardware and software manufacturers, established to develop standards for multimedia applications. The goal of the IMA in developing its audio compression standard was to have a public domain method that is simple and fast enough such that a 20-MHz 386-class personal computer would be able to decode, in real time, sound recorded in stereo at 44,100 16-bit samples per second (this is 88,200 16-bit samples per second). The encoder quantizes each 16-bit audio sample into a 4-bit code. The compression factor is therefore a constant 4. The “secret” of the IMA algorithm is the simplicity of its predictor. The predicted value Xp[n-1] that is output by the predictor is simply the decoded value Xp[n] of the preceding input X[n]. The predictor just stores Xp[n] for one cycle (one audio sample interval), then outputs it as Xp[n-1]. It does not use any of the preceding values Xp[i] to obtain better prediction. Thus, the predictor is not adaptive (but the quantizer is). Also, no side information is generated by the encoder. Figure 10.19a is a block diagram of the IMA quantizer. It is both simple and adaptive, varying the quantization step size based on both the current step size and the previous quantizer output. The adaptation is done by means of two table lookups, so it is fast. The quantizer outputs 4-bit codes where the leftmost bit is a sign and the remaining three bits are the number of quantization levels computed for the current audio sample. These three bits are used as an index to the first table. The item found in this table serves as an index adjustment to the second table. The index adjustment is added to a previously stored index, and the sum, after being checked for proper range, is used as the index for the second table lookup. The sum is then stored, and it becomes the stored index used in the next adaptation step. The item found in the second table becomes the new quantization step size. Figure 10.19b illustrates this process, and 10.7 MLP Audio 979 Tables 10.21 and 10.22 list the two tables. Table 10.20 shows the 4-bit output produced by the quantizer as a function of the sample size. For example, if the sample is in the range [1.5ss, 1.75ss), where ss is the step size, then the output is 0|110. Table 10.21 adjusts the index by bigger steps when the quantized magnitude is bigger. Table 10.22 is constructed such that the ratio between successive entries is about 1.1. ADPCM: Short for Adaptive Differential Pulse Code Modulation, a form of pulse code modulation (PCM) that produces a digital signal with a lower bit rate than standard PCM. ADPCM produces a lower bit rate by recording only the difference between samples and adjusting the coding scale dynamically to accommodate large and small differences. —From Webopedia.com 10.7 MLP Audio Note. The MLP audio compression method described in this section is different from and unrelated to the MLP (multilevel progressive) image compression method of Section 7.25. The identical acronyms are an unfortunate coincidence. Ambiguity refers to the property of words, terms, and concepts, (within a particular context) as being in undefined, undefinable, or otherwise vague, and thus having an unclear meaning. A word, phrase, sentence, or other communication is called “ambiguous” if it can be interpreted in more than one way. —From http://en.wikipedia.org/wiki/Disambiguation Meridian [Meridian 03] is a British company specializing in high-quality audio products, such as CD and DVD players, loudspeakers, radio tuners, and surround sound stereo amplifiers. Good-quality digitized sound, such as found in an audio CD, employs two channels (stereo sound), each sampled at 44.1 kHz with 16-bit samples (Section 10.2). A typical high-quality digitized sound, on the other hand, may use six channels (i.e., the sound is originally recorded by six microphones, for surround sound), sampled at the high rate of 96 kHz (to ensure that all the nuances of the performance can be delivered), with 24-bit samples (to get the highest possible dynamic range). This kind of audio data is represented by 6Χ96,000Χ24 = 13.824 Mbps (that’s megabits, not megabytes). In contrast, a DVD (digital versatile disc) holds 4.7 Gbytes, which at 13.824 Mbps in uncompressed form translates to only 45 minutes of playing. (Recall that even CDs, which have a much smaller capacity, hold 74 minutes of play time. This is an industry standard.) Also, the maximum data transfer rate for DVD-A (audio) is 9.6 Mbps, much lower than 13.824 Mbps. It is obvious that compression is the key to achieving a practical DVD-A format, but the high quality (as opposed to just good quality) requires lossless compression. The algorithm that has been selected as the compression standard for DVD-A (audio) is MLP (Meridian Lossless Packing). This algorithm is patented, and some of its details are still kept proprietary, which is reflected in the information provided in this 980 10. Audio Compression yes yes yes yes no no no no start sample<0 ? bit3‹0 bit2‹0 bit1‹0 bit0‹0 bit0‹1 bit3‹1 sample‹-sample sample. step size ? sample-step size sample‹ bit2‹1 lookup first table lookup second table [0,88] limit idex to next adaptation save index for sample-step size/2 sample‹ bit1‹1 sample. step size/4 ? sample. step size/2 ? done + adjust index new step ls 3 bits of quantizer output size (a) (b) Figure 10.19: (a) IMA ADPCM Quantization. (b) Step Size Adaptation. If sample 4-Bit If sample 4-Bit is in range quant is in range quant [1.75ss,.) 0|111 [-.,-1.75ss) 1|111 [1.5ss, 1.75ss) 0|110 [-1.75ss,-1.5ss) 1|110 [1.25ss, 1.5ss) 0|101 [-1.5ss,-1.25ss) 1|101 [1ss, 1.25ss) 0|100 [-1.25ss,-1ss) 1|100 [.75ss, 1ss) 0|011 [-1ss,-.75ss) 1|011 [.5ss, .75ss) 0|010 [-.75ss,-.5ss) 1|010 [.25ss, .5ss) 0|001 [-.5ss,-.25ss) 1|001 [0, .25ss) 0|000 [-.25ss, 0) 1|000 Table 10.20: Step Size and 4-Bit Quantizer Outputs. 10.7 MLP Audio 981 three bits quantized index magnitude adjust 000 -1 001 -1 010 -1 011 -1 100 2 101 4 110 6 111 8 Table 10.21: First Table for IMA ADPCM. Index Step Size Index Step Size Index Step Size Index Step Size 0 7 22 60 44 494 66 4,026 1 8 23 66 45 544 67 4,428 2 9 24 73 46 598 68 4,871 3 10 25 80 47 658 69 5,358 4 11 26 88 48 724 70 5,894 5 12 27 97 49 796 71 6,484 6 13 28 107 50 876 72 7,132 7 14 29 118 51 963 73 7,845 8 16 30 130 52 1,060 74 8,630 9 17 31 143 53 1,166 75 9,493 10 19 32 157 54 1,282 76 10,442 11 21 33 173 55 1,411 77 11,487 12 23 34 190 56 1,552 78 12,635 13 25 35 209 57 1,707 79 13,899 14 28 36 230 58 1,878 80 15,289 15 31 37 253 59 2,066 81 16,818 16 34 38 279 60 2,272 82 18,500 17 37 39 307 61 2,499 83 20,350 18 41 40 337 62 2,749 84 22,358 19 45 41 371 63 3,024 85 24,633 20 50 42 408 64 3,327 86 27,086 21 55 43 449 65 3,660 87 29,794 88 32,767 Table 10.22: Second Table for IMA ADPCM. 982 10. Audio Compression section. The term “packing” has a dual meaning. It refers to (1) removing redundancy from the original data in order to “pack” it densely, and (2) the audio samples are encoded in packets. MLP operates by reducing or completely removing redundancies in the digitized audio, without any quantization or other loss of data. Notice that high-quality audio formats such as 96 kHz with 24-bit samples carry more information than is strictly necessary for the human listener (or more than is available from modern microphone and converter techniques). Thus, such audio formats contain much redundancy and can be compressed efficiently. MLP can handle up to 63 audio channels and sampling rates of up to 192 kHz. The main features of MLP are as follows: 1. At least 4 bits/sample of compression for both average and peak data rates. 2. Easy transformation between fixed-rate and variable-rate data streams. 3. Careful and economical handling of mixed input sample rates. 4. Simple, fast decoding. 5. It is cascadable. An audio stream can be encoded and decoded multiple times in succession, and the output will always be an exact copy of the original. With MLP, there is no generation loss. The term “variable data rate” is important. An uncompressed data stream consists of audio samples, and each second of sound requires the same number of samples. Such a stream has a fixed data rate. In contrast, the compressed stream generated by a lossless audio encoder has a variable data rate; each second of sound occupies a different number of bits in this stream, depending on the nature of the sound. A second of silence occupies very few bits, whereas a second of random sound will not compress and will require the same number of bits in the compressed stream as in the original file. Most lossless audio compression methods are designed to reduce the average data rate, but MLP has the important feature that it reduces the instantaneous peak data rate by a known amount. This feature makes it possible to record 74 min of any kind of nonrandom sound on a 4.7-Gbyte DVD-A. In addition to being lossless (which means that the original data is delivered bit-forbit at the playback), MLP is also robust. It does not include any error-correcting code but has error-protection features. It uses check bits to make sure that each packet decompressed by the decoder is identical to that compressed by the encoder. The compressed stream contains restart points, placed at intervals of 10–30 ms. When the decoder notices an error, it simply skips to the next restart point, with a minimal loss of sound. This is another meaning of the term “high-quality sound.” For the first time, a listener hears exactly what the composer/performer intended—bit-for-bit and note-for-note. With lossy audio compression, the amount of compression is measured by the number of bits per second of sound in the compressed stream, regardless of the audio-sample size. With lossless compression, a large sample size (which really means more leastsignificant bits), must be losslessly compressed, so it increases the size of the compressed stream, but the extra LSBs typically have little redundancy and are therefore harder to compress. This is why lossless audio compression should be measured by the number of bits saved in each audio sample—a relative measure of compression. MLP reduces the audio samples from their original size (typically 24 bits) depending on the sampling rate. For average data rates, the reduction is as follows: For sampling 10.7 MLP Audio 983 rates of 44.1 kHz and 48 kHz, a sample is reduced by 5 to 11 bits. At 88.2 kHz and 96 kHz, the reduction increases to 9 to 13 bits. At 192 kHz, MLP can sometimes reduce a sample by 14 bits. Even more important are the savings for peak data rates. They are 4 bits for 44.1 kHz, 8 bits for 96 kHz, and 9 bits for 192 kHz samples. These peak data rate savings amount to a virtual guarantee, and they are one of the main reasons for the adoption of MLP as the DVD-A compression standard. The remainder of this section covers some of the details of MLP. It is based on [Stuart et al. 99]. The techniques used by MLP to compress audio samples include: 1. It looks for blocks of consecutive audio samples that are small, i.e., have several most-significant 0 bits. This is known as “dead air.” The LSBs of such samples are removed, which is equivalent to shifting the samples to the right. 2. It identifies channels that do not use their full bandwidth. 3. It identifies and removes correlation between channels. 4. It removes correlation between consecutive audio samples in each channel. 5. It uses buffering to smooth the production rate of the compressed output. The main compression steps are (1) lossless processing, (2) lossless matrixing, (3) lossless IIR filtering, (4) entropy encoding, and (5) FIFO buffering of the output. Lossless processing refers to identifying blocks of audio samples with unused capacity. Such samples are shifted. The term “matrixing” refers to reducing interchannel correlations by means of an affine transformation matrix. The IIR filter decorrelates the audio samples in each channel (intrachannel decorrelation) by predicting the next sample from its predecessors and producing a prediction difference. The differences are then compressed by an entropy encoder. The result is a compressed stream for each channel, and those streams are FIFO buffered to smooth the rate at which the encoded data is output. Finally, the output from the buffer is divided into packets, and check bits and restart points are added to each packet for error protection. An audio file with several channels is normally obtained by recording sound with several microphones placed at different locations. Each microphone outputs an analog signal that’s digitized and becomes an audio channel. Thus, the audio samples in the various channels are correlated, and reducing or removing this correlation is an important step, termed “matrixing,” in MLP compression. Denoting audio sample i of channel j by aij , matrixing is done by subtracting from aij a linear combination of audio samples aik from all other channels k. Thus, aij ‹ aij - k=j aikmkj, where column j of matrix m corresponds to audio channel j. This linear combination can also be viewed as a weighted sum where each audio sample aik is multiplied by a weight mkj. It can also be interpreted as an affine transformation of the vector of audio samples (ai,1, ai,2, . . . , ai,j-1, ai,j+1, . . . , ai,n). The next MLP encoding step is prediction. The matrixing step may remove all or part of the correlation between audio channels, but the individual samples in each channel are still correlated; a sample tends to be similar to its near neighbors, and the samples in an audio channel make up what is called quasi-periodic components. MLP uses linear prediction where a special IIR filter is applied to the data to remove the correlated 984 10. Audio Compression or timbral portion of the signal. The filter leaves a low-amplitude residual signal that is aperiodic and resembles noise. This residual signal represents sound components such as audio transients, phase relationships between partial harmonics, and even extra rosin on strings (no kidding). The entropy encoding step follows the matrixing (removing interchannel correlations) and prediction (removing intra-channel correlations). This step gets rid of any remaining correlations between the original audio samples. Several entropy encoders are built into the MLP encoder, and it may select any of them to encode any audio channel. Generally, MLP assumes that audio signals have a Laplacian distribution about zero, and its entropy encoders assign variable-length codes to the audio samples based on this assumption (recall that the MLP image compression method of Section 7.25 also uses this distribution, which is displayed graphically in Figure 7.143b). Buffering is the next step in MLP encoding. It smooths out the variations in bitrate caused by variations in the sound that’s being compressed. Passages of uniform sound or even silence may alternate with complex, random-like sound. When compressing such sound, the MLP encoder may output a few bits for each second of the uniform passages (which compress well) followed by many bits for each second for those parts of the sound that don’t compress well. Buffering ensures that the encoder outputs data to the compressed stream at a constant rate. This may not be very important when the compressed stream is a file, but it can make a big difference when the compressed audio is transmitted and is supposed to be received by an MLP decoder and played at real time without pauses. In addition to the compressed data, the compressed stream includes instructions to the decoder and CRC check bits. The input audio file is divided into blocks that are typically 40–160 samples long each. Blocks are assembled into packets whose length varies between 640 and 2,560 samples and is controlled by the user. Each packet is selfchecked and contains restart information. If the decoder senses bad data and cannot continue decompressing a packet, it simply skips to the start of the next packet, with a typical loss of just 7 ms of play time—hardly noticeable by a listener. MLP also has lossy options. One such option is to reduce all audio samples from 24 to 22 bits. Another is to pass some of the audio channels through a lowpass filter. This tends to reduce the entropy of the audio signal, making it possible for the MLP encoder to produce better compression. Such options may be important in cases where more than 74 min of audio should be recorded on one DVD-A. 10.8 Speech Compression Certain audio codecs are designed specifically to compress speech signals. Such signals are audio and are sampled like any other audio data, but because of the nature of human speech, they have properties that can be exploited for efficient compression. [Jayant 97] covers several types of speech codecs, but more information on this topic is available on the World Wide Web in researchers’ sites. This section starts by discussing the properties of human speech and continues with a description of various speech codecs. 10.8 Speech Compression 985 10.8.1 Properties of Speech We produce sound by forcing air from the lungs through the vocal cords into the vocal tract (Figure 10.23). The air ends up escaping through the mouth, but the sound is generated in the vocal tract (which extends from the vocal cords to the mouth, and in an average adult it is 17 cm long) by vibrations in much the same way as air passing through a flute generates sound. The pitch of the sound is controlled by varying the shape of the vocal tract (mostly by moving the tongue) and by moving the lips. The intensity (loudness) is controlled by varying the amount of air sent from the lungs. Humans are much slower than computers or other electronic devices, and this is also true with regard to speech. The lungs operate slowly, and the vocal tract changes shape slowly, so the pitch and loudness of speech vary slowly. When speech is captured by a microphone and is sampled, we find that adjacent samples are similar, and even samples separated by 20 ms are strongly correlated. This correlation is the basis of speech compression. from lungs Vocal cords Vocal tract tongue Nasal tract Lips Figure 10.23: A Cross Section of the Human Head. The vocal cords can open and close, and the opening between them is called the glottis. The movements of the glottis and vocal tract give rise to different types of sound. The three main types are as follows: 1. Voiced sounds. These are the sounds we make when we talk. The vocal cords vibrate, which opens and closes the glottis, thereby sending pulses of air at varying pressures to the tract, where it is shaped into sound waves. Varying the shape of the vocal cords and their tension changes the rate of vibration of the glottis and therefore controls the pitch of the sound. Recall that the ear is sensitive to sound frequencies of from 16 Hz to about 20,000–22,000 Hz. The frequencies of the human voice, on the other hand, are much more restricted and are generally in the range of 500 Hz to about 2 kHz. This is equivalent to time periods of 2 ms to 20 ms, and to a computer, such 986 10. Audio Compression periods are very long. Thus, voiced sounds have long-term periodicity, and this is the key to good speech compression. Figure 10.24a is a typical example of waveforms of voiced sound. 2. Unvoiced sounds. These are sounds that are emitted and can be heard, but are not parts of speech. Such a sound is the result of holding the glottis open and forcing air through a constriction in the vocal tract. When an unvoiced sound is sampled, the samples show little correlation and are random or close to random. Figure 10.24b is a typical example of the waveforms of unvoiced sound. 3. Plosive sounds. These result when the glottis closes, the lungs apply air pressure on it, and it suddenly opens, letting the air escape suddenly. The result is a popping sound. There are also combinations of the above three types. An example is the case where the vocal cords vibrate and there is also a constriction in the vocal tract. Such a sound is called fricative. 0 5 10 15 20 25 30 Time (ms) 4000 3000 2000 1000 0 0 1000 2000 3000 4000 800 600 400 200 200 400 600 800 Amplitude 0 5 10 15 20 25 30 (a) (b) Time (ms) Figure 10.24: (a) Voiced and (b) Unvoiced Sound Waves. Speech codecs. There are three main types of speech codecs. Waveform speech codecs produce good to excellent speech after compressing and decompressing it, but generate bitrates of 10–64 kbps. Source codecs (also called vocoders) generally produce poor to fair speech but can compress it to very low bitrates (down to 2 kbps). Hybrid codecs are combinations of the former two types and produce speech that varies from fair to good, with bitrates between 2 and 16 kbps. Figure 10.25 illustrates the speech quality versus bitrate of these three types. 10.8.2 Waveform Codecs This type of codec does not attempt to predict how the original sound was generated. It only tries to produce, after decompression, audio samples that are as close to the original ones as possible. Thus, such codecs are not designed specifically for speech coding and can perform equally well on all kinds of audio data. As Figure 10.25 illustrates, when such a codec is forced to compress sound to less than 16 kbps, the quality of the reconstructed sound drops significantly. The simplest waveform encoder is pulse code modulation (PCM). This encoder simply quantizes each audio sample. Speech is typically sampled at only 8 kHz. If 10.8 Speech Compression 987 1 2 4 Speech quality good 8 16 32 64 codecs codecs Hybrid Vocoders Bitrate (kbps) bad poor fair Waveform Figure 10.25: Speech Quality Versus Bitrate for Speech Codecs. each sample is quantized to 12 bits, the resulting bitrate is 8k Χ 12 = 96 kbps and the reproduced speech sounds almost natural. Better results are obtained with a logarithmic quantizer, such as the ΅-law and A-law companding methods (Section 10.5). They quantize audio samples to varying numbers of bits and may compress speech to 8 bits per sample on average, thereby resulting in a bitrate of 64 kbps, with very good quality of the reconstructed speech. A differential PCM (DPCM; Section 10.6) speech encoder uses the fact that the audio samples of voiced speech are correlated. This type of encoder computes the difference between the current sample and its predecessor and quantizes the difference. An adaptive version (ADPCM; Section 10.6) may compress speech at good quality down to a bitrate of 32 kbps. Waveform coders may also operate in the frequency domain. The subband coding algorithm (SBC) transforms the audio samples to the frequency domain, partitions the resulting coefficients into several critical bands (or frequency subbands), and codes each subband separately with ADPCM or a similar quantization method. The SBC decoder decodes the frequency coefficients, recombines them, and performs the inverse transformation to (lossily) reconstruct audio samples. The advantage of SBC is that the ear is sensitive to certain frequencies and less sensitive to others (Section 10.3, especially Table 10.6). Subbands of frequencies to which the ear is less sensitive can therefore be coarsely quantized without loss of sound quality. This type of coder typically produces good reconstructed speech quality at bitrates of 16–32 kbps. They are, however, more complex to implement than PCM codecs and may also be slower. The adaptive transform coding (ATC) speech compression algorithm transforms audio samples to the frequency domain with the discrete cosine transform (DCT). The audio file is divided into blocks of audio samples and the DCT is applied to each block, resulting in a number of frequency coefficients. Each coefficient is quantized according to the frequency to which it corresponds. Good quality reconstructed speech can be achieved at bitrates as low as 16 kbps. 988 10. Audio Compression 10.8.3 Source Codecs In general, a source encoder uses a mathematical model of the source of data. The model depends on certain parameters, and the encoder uses the input data to compute those parameters. Once the parameters are obtained, they are written (after being suitably encoded) on the compressed stream. The decoder inputs the parameters and employs the mathematical model to reconstruct the original data. If the original data is audio, the source coder is called vocoder (from “vocal coder”). The vocoder described in this section is the linear predictive coder (LPC, see [Rabiner and Schafer 78]). Figure 10.26 shows a simplified model of speech production. Part (a) illustrates the process in a person, whereas part (b) shows the corresponding LPC mathematical model. Voiced waveforms Unvoiced waveforms Vocal Vocal cords tract Lip control Speech output Voiced amplitude Unvoiced T=pitch period periodic samples White noise Innovations LPC filter Speech amplitude Noise source Excitation sources Pitch sources samples u(n) s(n) H(z) G V UV (a) (b) Figure 10.26: Speech Production: (a) Real. (b) LPC Model. In this model, the output is the sequence of speech samples s(n) coming out of the LPC filter (which corresponds to the vocal tract and lips). The input u(n) to the model (and to the filter) is either a train of pulses (when the sound is voiced speech) or white noise (when the sound is unvoiced speech). The quantities u(n) are also termed innovation. The model illustrates how samples s(n) of speech can be generated by mixing innovations (a train of pulses and white noise). Thus, it represents mathematically the relation between speech samples and innovations. The task of the speech encoder is to input samples s(n) of actual speech, use the filter as a mathematical function to determine an equivalent sequence of innovations u(n), and output the innovations in compressed form. The correspondence between the model’s parameters and the parts of real speech is as follows: 10.8 Speech Compression 989 1. Parameter V (voiced) corresponds to the vibrations of the vocal cords. UV expresses the unvoiced sounds. 2. T is the period of the vocal cords vibrations. 3. G (gain) corresponds to the loudness or the air volume sent from the lungs each second. 4. The innovations u(n) correspond to the air passing through the vocal tract. 5. The symbols . and . denote amplification and combination, respectively. The main equation of the LPC model describes the output of the LPC filter as H(z) = 1 1 + a1z-1 + a2z-2 + · · · + a10z-10 , where z is the input to the filter [the value of one of the u(n)]. An equivalent equation describes the relation between the innovations u(n) on the one hand and the 10 coefficients ai and the speech audio samples s(n) on the other hand. The relation is u(n) = s(n) + 10 i=1 ais(n - i). (10.4) This relation implies that each number u(n) input to the LPC filter is the sum of the current audio sample s(n) and a weighted sum of the 10 preceding samples. The LPC model can be written as the 13-tuple A = (a1, a2, . . . , a10, G, V/UV, T), where V/UV is a single bit specifying the source (voiced or unvoiced) of the input samples. The model assumes that A stays stable for about 20 ms, then gets updated by the audio samples of the next 20 ms. At a sampling rate of 8 kHz, there are 160 audio samples s(n) every 20 ms. The model computes the 13 quantities in A from these 160 samples, writes A (as 13 numbers) on the compressed stream, then repeats for the next 20 ms. The resulting compression factor is therefore 13 numbers for each set of 160 audio samples. It’s important to distinguish the operation of the encoder from the diagram of the LPC’s mathematical model depicted in Figure 10.26b. The figure shows how a sequence of innovations u(n) generates speech samples s(n). The encoder, however, starts with the speech samples. It inputs a 20-ms sequence of speech samples s(n), computes an equivalent sequence of innovations, compresses them to 13 numbers, and outputs the numbers after further encoding them. This repeats every 20 ms. LPC encoding (or analysis) starts with 160 sound samples and computes the 10 LPC parameters ai by minimizing the energy of the innovation u(n). The energy is the function f(a1, a2, . . . , a10) = 159 i=0 u2(n) 990 10. Audio Compression and its minimum is computed by differentiating it 10 times, with respect to each of its 10 parameters and setting the derivatives to zero. The 10 equations .f .ai = 0, for i = 1, 2, . . . , 10 can be written in compact notation ............... R(0) R(1) R(2) R(3) R(4) R(5) R(6) R(7) R(8) R(9) R(1) R(0) R(1) R(2) R(3) R(4) R(5) R(6) R(7) R(8) R(2) R(1) R(0) R(1) R(2) R(3) R(4) R(5) R(6) R(7) R(3) R(2) R(1) R(0) R(1) R(2) R(3) R(4) R(5) R(6) R(4) R(3) R(2) R(1) R(0) R(1) R(2) R(3) R(4) R(5) R(5) R(4) R(3) R(2) R(1) R(0) R(1) R(2) R(3) R(4) R(6) R(5) R(4) R(3) R(2) R(1) R(0) R(1) R(2) R(3) R(7) R(6) R(5) R(4) R(3) R(2) R(1) R(0) R(1) R(2) R(8) R(7) R(6) R(5) R(4) R(3) R(2) R(1) R(0) R(1) R(9) R(8) R(7) R(6) R(5) R(4) R(3) R(2) R(1) R(0) ............... ............... a1 a2 a3 a4 a5 a6 a7 a8 a9 a10 ............... = ............... -R(1) -R(2) -R(3) -R(4) -R(5) -R(6) -R(7) -R(8) -R(9) -R(10) ............... , where R(k), which is defined as R(k) = 159-k n=0 s(n)s(n + k), is the autocorrelation function of the samples s(n). This system of 10 algebraic equations is easy to solve numerically, and the solutions are the 10 LPC parameters ai. The remaining three parameters, V/UV , G, and T, are determined from the 160 audio samples. If those samples exhibit periodicity, then T becomes that period and the 1-bit parameter V/UV is set to V . If the 160 samples do not feature any well-defined period, then T remains undefined and V/UV is set to UV . The value of G is determined by the largest sample. LPC decoding (or synthesis) starts with a set of 13 LPC parameters and computes 160 audio samples as the output of the LPC filter by H(z) = 1 1 + a1z-1 + a2z-2 + · · · + a10z-10 . These samples are played at 8,000 samples per second and result in 20 ms of (voiced or unvoiced) reconstructed speech. The 2.4-kbps version of the LPC vocoder goes through one more step to encode the 10 parameters ai. It converts them to 10-line spectrum pair (LSP) parameters, denoted by .i, by means of the relations P(z) = 1+(a1 - a10)z-1 + (a2 - a9)z-2 + · · · + (a10 - a1)z-10 - z-11 = (1 - z-1).k=2,4,...,10(1 - 2 cos .kz-1 + z-2), Q(z) = 1+(a1 + a10)z-1 + (a2 + a9)z-2 + · · · + (a10 + a1)z-10 + z-11 = (1+z-1).k=1,3,...,9(1 - 2 cos .kz-1 + z-2). 10.8 Speech Compression 991 The ten LSP parameters satisfy 0 < .1 < .2 < · · · < .10 < .. Each is quantized to either three or four bits, as shown here, for a total of 34 bits. .1 .2 .3 .4 .5 .6 .7 .8 .9 .10 3 4 4 4 4 3 3 3 3 3 The gain parameter G is encoded in seven bits, the period T is quantized to six bits, and the V/UV parameter requires just one bit. Thus, the 13 LPC parameters are quantized to 48 bits and provide enough data for a 20-ms frame of decompressed speech. Each second of speech has 50 frames, so the bitrate is 50Χ48 = 2.4 kbps. Assuming that the original speech samples were eight bits each, the compression factor is (8,000Χ8)/2,400 = 26.67—very impressive. 10.8.4 Hybrid Codecs This type of speech codec combines features from both waveform and source codecs. The most popular hybrid codecs are Analysis-by-Synthesis (AbS) time-domain algorithms. Like the LPC vocoder, these codecs model the vocal tract by a linear prediction filter, but use an excitation signal instead of the simple, two-state voice-unvoice model to supply the u(n) (innovation) input to the filter. Thus, an AbS encoder starts with a set of speech samples (a frame), encodes them similar to LPC, decodes them, and subtracts the decoded samples from the original ones. The differences are sent through an error minimization process that outputs improved encoded samples. These samples are again decoded, subtracted from the original samples, and new differences computed. This is repeated until the differences satisfy a termination condition. The encoder then proceeds to the next set of speech samples (next frame). One of the best-known AbS codecs is CELP, an acronym that stands for codeexcited linear predictive. The 4.8-kbps CELP codec is similar to the LPC vocoder, with the following differences: 1. The frame size is 30 ms (i.e., 240 speech samples per frame and 33.3 frames per second of speech). 2. The innovation u(n) is coded directly. 3. A pitch prediction is included in the algorithm. 4. Vector quantization is used. Each frame is encoded to 144 bits as follows: The LSP parameters are quantized to a total of 34 bits. The pitch prediction filter becomes a 48-bit number. Codebook indices consume 36 bits. Four gain parameters require 5 bits each. One bit is used for synchronization, four bits encode the forward error control (FEC, similar to CRC), and one bit is reserved for future use. At 33.3 frames per second, the bitrate is therefore 33.3 Χ 144 = 4,795 . 4.8 kbps. There is also a conjugate-structured algebraic CELP (CS-ACELP) that uses 10 ms frames (i.e., 80 samples per frame) and encodes the LSP parameters with a two-stage vector quantizer. The gains (two per frame) are also encoded with vector quantization. 992 10. Audio Compression 10.9 Shorten Shorten is a simple, special-purpose, lossless compressor for waveform files. Any file whose data items (which are referred to as samples) go up and down as in a wave can be efficiently compressed by this method. Its developer [Robinson 94] had in mind applications to speech compression (where audio files with speech are distributed on a CD-ROM), but any other waveform files can be similarly compressed. The compression performance of Shorten isn’t as good as that of mp3, but Shorten is lossless. Shorten performs best on files with low-amplitude and low-frequency samples, where it yields compression factors of 2 or better. It has been implemented on UNIX and on MS-DOS and is freely available at [Softsound 03]. Shorten encodes the individual samples of the input file by partitioning the file into blocks, predicting each sample from some of its predecessors, subtracting the prediction from the sample, and encoding the difference with a special variable-length code. It also has a lossy mode, where samples are quantized before being compressed. The algorithm has been implemented by its developer and can input files in the audio formats ulaw (Section 10.5), s8, u8, s16 (this is the default input format), u16, s16x, u16x, s16hl, u16hl, s16lh, and u16lh, where “s” and “u” stand for “signed” and “unsigned,” respectively, a trailing “x” specifies byte-mapped data, “hl” implies that the high-order byte of a sample is followed in the file by the low-order byte, and “lh” signifies the reverse. An entire file is encoded in blocks, where the block size (typically 128 or 256 samples) is specified by the user and has a default value of 256. (At a sampling rate of 16 kHz, this is equivalent to 16 ms of sound.) The samples in the block are first converted to integers with an expected mean of 0. The idea is that the samples within each block have the same spectral characteristic and can therefore be predicted accurately. Some audio files consist of several interleaved channels, so Shorten starts by separating the channels in each block. Thus, if the file has two channels and the samples are interleaved as L1, R1, L2, R2, and so on up to Lb, Rb, the first step creates the two blocks (L1, L2, . . . , Lb) and (R1,R2, . . . , Rb) and each block is then compressed separately. In practice, blocks that correspond to audio channels are often highly correlated, so sophisticated methods, such as MLP (Section 10.7), try to remove interblock correlations before tackling the samples within a block. Once a block has been constructed, its samples are predicted and difference values are computed. A predicted value ˆs(t) for the current sample s(t) is computed from the p immediately-preceding samples by a linear combination (see also Section 7.30) ˆs(t) = p i=1 ais(t - i). If the prediction is done properly, then the difference (also termed error or residual) e(t) = s(t) - ˆs(t) will almost always be a small (positive or negative) number. The simplest type of wave is stationary. In such a wave, a single set of coefficients ai always produces the best prediction. Naturally, most waves are not stationary and should select a different set of ai coefficients to predict each sample. Such selection can be done in different ways, involving more and more neighbor samples, and this results in predictors of different orders. 10.9 Shorten 993 A zeroth-order predictor simply predicts each sample s(t) as zero. A first-order predictor (Figure 10.27a) predicts each s(t) as its predecessor s(t - 1). Similarly, a second-order predictor (Figure 10.27b) computes a straight segment (a linear function or a degree-1 polynomial) from s(t - 2) to s(t - 1) and continues it to predict s(t). Extending this idea to one more point, a third-order predictor (Figure 10.27c) computes a degree-2 polynomial (a conic section) that passes through the three points s(t - 3), s(t-2), and s(t-1) and extrapolates it to predict s(t). In general, an nth-order predictor computes a degree-(n-1) polynomial that passes through the n points s(t-n) through s(t - 1) and extrapolates it to predict s(t). We show how to compute second- and third-order predictors. (a) (b) (c) • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • Figure 10.27: Predictors of Orders 1, 2, and 3. Given the two points P2 = (t-2, s2) and P1 = (t-1, s1), we can write the parametric equation of the straight segment connecting them as L(u) = (1-u)P2 +uP1 = (1-u)(t-2, s2)+u(t-1, s1) = (u+t-2, (1-u)s2 +us1). It’s easy to see that L(0) = P2 and L(1) = P1. Extrapolating to the next point, at u = 2, yields L(2) = (t, 2s1-s2). Using our notation, we conclude that the second-order predictor predicts sample s(t) as the linear combination 2s(t - 1) - s(t - 2). For the third-order predictor, we start with the three points P3 = (t - 3, s3), P2 = (t - 2, s2), and P1 = (t - 1, s1). The degree-2 polynomial that passes through those points is given by the uniform quadratic Lagrange interpolation polynomial (see, for example, [Salomon 06] p. 78, Equation 3.12.) L(u) = [u2, u, 1].. 1 2 -1 1 2 -3 2 2 -1 2 1 0 0 .. .. P3 P2 P1 .. = u2 2 - 3u 2 + 1P3 + (-u2 + 2u)P2 + u2 2 - u 2 P1. It is easy to verify that L(0) = P3, L(1) = P2, and L(2) = P1. Extrapolating to u = 3 yields L(3) = 3P1 - 3P2 + P3. When this is translated to our samples, the result is 994 10. Audio Compression ˆs(t) = 3s(t - 1) - 3s(t - 2) + s(t - 3). The first four predictors are summarized as ˆs0(t) = 0, ˆs1(t) = s(t - 1), ˆs2(t) = 2s(t - 1) - s(t - 2), ˆs3(t) = 3s(t - 1) - 3s(t - 2) + s(t - 3). (10.5) These predictors can now be used to compute the error (or difference) values for the first four orders: e0(t) = s(t) - ˆs0(t) = s(t), e1(t) = s(t) - ˆs1(t) = s(t) - s(t - 1) = e0(t) - e0(t - 1), e2(t) = s(t) - ˆs2(t) = s(t) - 2s(t - 1) + s(t - 2) = [s(t) - s(t - 1)] - [s(t - 1) - s(t - 2)] = e1(t) - e1(t - 1), (10.6) e3(t) = s(t) - ˆs3(t) = s(t) - 3s(t - 1) + 3s(t - 2) - s(t - 3) = [s(t) - 2s(t - 1) + s(t - 2)] - [s(t - 1) - 2s(t - 2) + s(t - 3)] = e2(t) - e2(t - 1). This computation is recursive but it involves only three steps, it is arithmetically simple, and does not require any multiplications (see also Equation (10.8)). For maximum compression, it is possible to compute all four predictors and their errors and select the smallest error. However, experience gained by the developer of the method indicates that even a zeroth-order predictor results in typical compression of 48%, and going all the way to third-order prediction improves this only to 58%. For most cases, there is therefore no need to use higher-order predictors, and the precise predictor used should be determined by compression quality versus run time considerations. The default mode of Shorten uses linear (second-order) prediction. The error (or difference) values are assumed decorrelated and are replaced by variable-length codes. Strong correlation between the original samples implies that most error values are small or even zero, and only few are large. They are also signed. Experiments suggest that the distribution of the errors is very close to the well-known Laplace distribution (Figure 7.143b), suggesting that the variable-length codes used to encode the differences should follow this distribution. (The developer of this method claims that the error introduced by coding the difference values according to the Laplace distribution, instead of using the actual probability distribution of the differences, is only 0.004 bits.) The codes selected for Shorten have been developed by Robert F. Rice and are known as the Rice codes ([Rice 79], [Rice 91], and [Fenwick 96a]). They are a special case of the Golomb code (Section 3.24) and are also related to the subexponential code of Section 7.24.1. A Rice code depends on the choice of a base n and is computed in the following steps: (1) Separate the sign bit from the rest of the number. This becomes the most-significant bit of the Rice code. (2) Separate the n LSBs. They become the LSBs of the Rice code. (3) Code the remaining bits in unary and make this the middle part of the Rice code. (If the remaining bits are, say, 11, then the unary code is either 10.9 Shorten 995 three zeros followed by a 1 or three 1’s followed by a 0.) Thus, this code is computed with a few logical operations, which is faster than computing a Huffman code, which requires sliding down the Huffman tree while collecting the individual bits of the code. This feature is especially important for the decoder, which has to be simple and fast. Table 10.28 shows examples of this code for n = 2 (the column labeled “No. of zeros” lists the number of zeros in the unary part of the code). No. of No. of i Binary Sign LSB Zeros Code i Binary Sign LSB Zeros Code 0 0 0 00 0 0| 1|00 1 1 0 01 0 0| 1|01 -1 1 1 01 0 1| 1|01 2 10 0 10 0 0| 1|10 -2 10 1 10 0 1| 1|10 3 11 0 11 0 0| 1|11 -3 11 1 11 0 1| 1|11 4 100 0 00 1 0|01|00 -4 100 1 00 1 1|01|00 5 101 0 01 1 0|01|01 -5 101 1 01 1 1|01|01 6 110 0 10 1 0|01|10 -6 110 1 10 1 1|01|10 7 111 0 11 1 0|01|11 -7 111 1 11 1 1|01|11 8 1000 0 00 2 0|001|00 -8 1000 1 00 2 1|001|00 11 1011 0 11 2 0|001|11 -11 1011 1 11 2 1|001|11 12 1100 0 00 3 0|0001|00 -12 1100 1 00 3 1|0001|00 15 1111 0 11 3 0|0001|11 -15 1111 1 11 3 1|0001|11 Table 10.28: Various Positive and Negative Rice Codes. The Rice code, similar to the Huffman code, assigns an integer number of bits to each difference value, so in general, it is not an entropy code. However, the developer of this method claims that the error introduced by using these codes is only 0.12 bits per difference value. Rice codes are ideal for data items with a Laplace distribution, but other prefix codes exist that are easier to construct and to decode and that may, under certain circumstances, outperform the Rice codes. Table 10.29 lists three such codes. The “pod” code, due to Robin Whittle [firstpr 06], codes the number zero with the single bit 1, and codes the binary number 1 b...b 3456 k as 0...0 3456 k+1 1 b...b 3456 k . In two cases, the pod code is one bit longer than the Rice code, in four cases it has the same length, and in all other cases it is shorter than the Rice codes. The Elias Gamma code [Fenwick 96a] is identical to the pod code minus its leftmost zero. It is therefore shorter, but does not include a code for zero. The biased Elias Gamma code corrects this fault in an obvious way but at the cost of making some codes one bit longer. There remains the question of what base value n to select for the Rice codes. The base determines how many low-order bits of a difference value are included directly in the Rice code, and this is linearly related to the variance of the difference values. The developer provides the formula n = log2[log(2)E(|x|)], where E(|x|) is the expected value of the differences. This value is the sum |x|p(x) taken over all possible differences x. 996 10. Audio Compression Number Pod Elias Biased Elias Dec Binary Gamma Gamma 0 00000 1 1 1 00001 01 1 010 2 00010 0010 010 011 3 00011 0011 011 00100 4 00100 000100 00100 00101 5 00101 000101 00101 00110 6 00110 000110 00110 00111 7 00111 000111 00111 0001000 8 01000 00001000 0001000 0001001 9 01001 00001001 0001001 0001010 10 01010 00001010 0001010 0001011 11 01011 00001011 0001011 0001100 12 01100 00001100 0001100 0001101 13 01101 00001101 0001101 0001110 14 01110 00001110 0001110 0001111 15 01111 00001111 0001111 000010000 16 10000 0000010000 000010000 000010001 17 10001 0000010001 000010001 000010010 18 10010 0000010010 000010010 000010011 Table 10.29: Pod, Elias Gamma, and Biased Elias Gamma Codes. The steps described so far result in lossless coding of the audio samples. Sometimes, a certain loss in the samples is acceptable if it results in significantly better compression. Shorten offers two options of lossy compression by quantizing the original audio samples before they are compressed. The quantization is done separately for each segment. One lossy option encodes every segment at the same bitrate (the user specifies the maximum bitrate), and the other option maintains a user-specified signal-to-noise ratio in each segment (the user specifies the minimum acceptable signal-to-noise ratio in dB). Tests of Shorten indicate that it generally performs better and faster than UNIX compress and gzip, but that the lossy options are slow. 10.10 FLAC The name FLAC is an acronym that stands for free lossless audio compression. FLAC was especially designed for audio compression and it also supports streaming and archival of audio data. FLAC is the brainchild of Josh Coalson who developed it in 1999 based on ideas from Shorten. He then started the FLAC project on the well-known sourceforge Web site [sourceforge.flac 06] by releasing his reference implementation. Since then many developers have contributed to improving the reference implementation and to writing alternative implementations. The FLAC project, administered and coordinated by Josh Coalson, maintains the software and provides a reference codec and input plugins for several popular audio players. 10.10 FLAC 997 FLAC is free both in the sense that it is available at no cost and that its specifications and format are fully open to the public and can be used for any purpose. This includes the FLAC source code, which is available under open-source licenses. Neither the FLAC format nor any of the implemented encoding/decoding algorithms are covered by any known patents. Thus, FLAC is one of the first open and free lossless audio compression methods. It is possible that Squish (or Ogg Squish, by Chris Montgomery, [Ogg squish 06]), holds the title of “first,” though it is not widely used.] The FLAC project consists of (1) the format of the compressed stream, (2) reference encoders and decoders in library form, (3) a command-line software FLAC codec, (4) metaflac, a command-line metadata editor for FLAC files, and (5) input plugins for various music players. The reference implementation of FLAC compiles on many platforms and supports many operating systems. Examples are Windows, BeOS, OS/2, and most Unix and Unix-like systems (including Linux, *BSD, Solaris, and Mac OS X). SourceForge.net is the world’s largest Open Source software development web site, hosting more than 100,000 projects and over 1,000,000 registered users with a centralized resource for managing projects, issues, communications, and code. Source- Forge.net has the largest repository of Open Source code and applications available on the Internet, and hosts more Open Source development products than any other site or network worldwide. SourceForge.net provides a wide variety of services to projects we host, and to the Open Source community. SourceForge.net provides free hosting to Open Source software development projects. The essence of the Open Source development model is the rapid creation of solutions within an open, collaborative environment. Collaboration within the Open Source community (developers and end users) promotes a higher standard of quality, and helps to ensure the long-term viability of both data and applications. SourceForge.net is owned by the Open Source Technology Group (OSTG). From http://sourceforge.net/docs/about FLAC compresses the audio input block by block and is based on prediction and Rice codes. Thus, it somewhat resembles Shorten (Section 10.9). A block of audio samples is compressed by predicting the samples, subtracting each audio sample from its prediction, and encoding the difference with a Rice code (Sections 3.24 and 10.9). The compressed stream consists of the Rice codes, a few parameters specifying the prediction, and metadata information. The latter includes information such as the audio sampling rate, the number of audio channels, the minimum and maximum data rates, the minimum and maximum block sizes, the MD5 signature of the unencoded audio data, padding, seek tables, tags, cuesheets, and application-specific data. Users who need custom metadata can define their own format and request a metadata ID from the FLAC developers at [flacID 06]. FLAC does not forbid copy-prevention schemes such as DRM from appearing in FLAC streams, it just does not make any provisions for copy prevention in the format. It is possible, say, for Apple to encrypt a FLAC stream with FairPlay and store it in an m4a container. Many audiophiles who are concerned about high audio quality are suspicious of lossy compression methods such as mp3 or AAC, and it is true that lossy methods are sensitive 998 10. Audio Compression to certain types of sound and may sometimes produce reconstructed sound of inferior quality. Such users/listeners prefer FLAC, and they have made this format popular. As a result, audio equipment manufacturers have started supporting this format in hardware. Currently (mid 2006), many home stereo, car stereo, portable, and handheld audio devices can play music in this format. Software is also available on many computer platforms and on most operating systems—including Windows, Unix (Linux, *BSD, Solaris, OS X, IRIX), BeOS, OS/2, and Amiga—to encode and decode sound in FLAC. A list of available devices and software is maintained at [flac.devices 06]. Because of its lossless nature, FLAC is limited to only integer audio samples. There are audio encoders that can input audio samples in floating-point, but floating-point computations often return slightly different results on different computers, because of differences in word size and in the way ALU circuits handle rounding errors. Restricting FLAC to integer audio samples ensures that a correct FLAC implementation will be able to fully and truly reconstruct any audio file. FLAC can handle audio sampling rates over a wide range. Any sampling rates of 1 Hz to 1,048,570 Hz (in 1-Hz increments) can be input and compressed correctly. FLAC is designed to handle up to eight audio channels (that can be grouped in stereo or as in 5.1 channel surround, page 1083) and take advantage of interchannel correlations to improve compression. FLAC Goals. Being open source means that anyone can volunteer to help in the development of FLAC. Thus, it is important to have a well-defined set of goals for this project. Developers should keep these goals in mind when trying to improve, modify, or maintain any part of FLAC, but the project administrator may modify the goals from time to time, to reflect changes in thinking and users’ needs. The current goals are: FLAC is and should stay an open format with an open-source reference implementation. FLAC is and should stay lossless. Future developers should not add lossy compression even as an option. The compression efficiency of FLAC should be on par or better than other lossless codecs. Decoding should be fast. A FLAC decoder should work at least in real time (i.e., decode the audio while it is played) even on average-speed hardware. FLAC should support fast sample-accurate seeking. A user listening to decoded audio should be able to skip to any point in the audio with ease. Being lossless, FLAC should allow for the playback of consecutive audio streams without gaps. No copy protection of any kind is allowed in a FLAC compressed stream. These goals imply the main features of FLAC, which are the following: Lossless. There are many excellent lossy audio compression methods, which is why FLAC’s developers decided to design it as lossless and keep it lossless. The decoded audio should be identical to the original input, and each implementation has to be tested to verify this property. FLAC has a “verify” option (-V) to verify the output while encoding. 10.10 FLAC 999 The decoder runs in parallel with the encoder, and the decoder’s output is compared with the original input. The software signals an error if a mismatch is found. Each frame in the compressed stream includes a 16-bit CRC of the frame data for detecting (but not correcting) transmission errors. It has already been mentioned that the header of the compressed file contains the MD5 signature of the original (uncompressed) data. This can be used by the decoder to verify its output and can also be employed by users and implementors to test an implementation. Fast. FLAC was designed for fast decoding. This is important in audio compression because the decoder often has to output the audio in real time, to be decoded and played simultaneously. FLAC is therefore an asymmetric compression method. Decoding is fast because it is simple (much less computationally intensive than perceptual decoders) and uses only integer operations. Hardware support: Because of FLAC’s free reference implementation and low decoding complexity, FLAC is currently the only lossless audio codec that has any kind of hardware support. Streamable. The FLAC compressed stream consists of frames, where each frame contains all the data necessary to decode the frame. Decoding a frame does not require any previous frames. The FLAC compressed stream includes sync codes and CRCs (similar to MPEG and other formats) that make it easy for the decoder to quickly skip to any point in the compressed stream. Seekable. Being able to skip to any point in the audio is useful for playback as well as for audio editing applications. Flexible metadata. The FLAC compressed stream starts with metadata blocks that include useful information. Users can design their own metadata blocks, for special and private information, and can apply to the FLAC project for block IDs. Suitable for archiving. An archive can benefit from compressed data, but it requires lossless compression. Lossless compression is also handy for converting between formats. The fact that FLAC is open source also encourages users to select FLAC for archiving audio. Convenient CD archiving. One of the metadata blocks on the FLAC compressed stream has a “cue sheet” for storing a CD table-of-contents and all track and index points. In a typical application, a CD is converted to a single file and its cue sheet is extracted. The file is then compressed into a single FLAC file with a metadata block containing the extracted cue sheet. If the original CD is damaged, the FLAC file with the cue sheet can be converted back and burned to become an exact copy of the CD. Error resistant. Many lossless codecs do not include any features for data reliability or integrity. A single error in a compressed stream may prevent the decoder from continuing. The FLAC compressed stream, in contrast, consists of frames, each representing a small fraction of a second of audio. If an error damages a frame, the decoder can still locate the next frame and continue. This limits any damages caused to the audio by errors during transmission or storage; an important, practical feature. Before we get to the details of the FLAC algorithm, here are the definitions of the two key concepts, block and frame. A block in FLAC is a number of consecutive audio samples. If the input has more than one audio channel, a block consists of one subblock 1000 10. Audio Compression for each channel. Thus, a subblock is a set of contiguous audio samples from the same channel. The block size is actually the size of a subblock. Thus, if the block size is 4,608 and there are two audio channels, then the actual size of the block is 4,608 Χ 2 = 9,216 samples. The block size is important, and it affects the efficiency, speed, and reliability of a FLAC codec. A small block represents a very small fraction of a second of audio. In case of an error during decoding, only that small period of time is lost. On the downside, small blocks imply many blocks, which presents more work for both the encoder and decoder. Also, each block becomes a frame in the compressed stream and a frame has a header. A large number of headers results in more overhead which leads to less compression. Another consideration is that large blocks feature reduced correlation between their audio samples and thus lead to inefficient prediction and compression. In a typical case where the sampling rate is 44.1 kHz and linear prediction is used, the optimal block size is 2,000–6,000 samples and the default value in such a case is 4,608 samples (except when the user-controlled option -l is set to zero, in which case the block size becomes 1,152). A frame is written on the FLAC output file for each block of the input file. The frame starts with a header which is followed by a number of subframes. Each subframe starts with its own header, followed by Rice codes for encoded audio samples from the same channel. A frame consists of one subframe for each audio channel, and each subframe consists of the same number of (Rice encoded) audio samples. A FLAC output stream starts with the four-byte identifying string fLaC. This is followed by the STREAMINFO metadata block. Other metadata blocks may optionally follow. They include all the side information that the decoder and end-user need. Metadata blocks can be of any length, and new ones can be defined by users. Thus, a FLAC decoder may not recognize certain metadata blocks and is allowed to skip them silently. The remainder of the stream consists of frames (there must be at least one such frame). The encoder. The FLAC encoder encodes each block of audio samples separately. The first step is to optionally exploit the inter-channel correlation of audio samples. If the input is stereo (left and right channels), it is decorrelated by means of the usual lossless transformation mid = (left + right)/2 and side = left - right. FLAC has two user-selected options for this decorrelation. Option -m directs the encoder to encode both the left and right channels and the mid and side channels, and then select the better (i.e., smaller) result. Option -M tells the encoder to switch between left-right and mid-side adaptively and always select the smaller frame. The next encoder step is modeling. Given a block of audio samples ai (from the left, right, mid, side, or any other channel), the encoder tries to find a function f(t) such that f(ti) is close to ai for all values of i and for certain values ti. The function (termed the approximation function or the fit) is fully specified by a few parameters which are sent to the decoder as side information. The encoder then subtracts ai -f(ti) to obtain a difference (or residue) ei. If the function closely fits the audio samples, the residues will be small integers (and may, of course, be negative). FLAC employs four methods for constructing an approximating function (although only the last two are general and provide compression). Verbatim. This method always predicts a zero audio sample. The difference (or residue) is therefore the original audio sample. It is pointless to replace the audio samples 10.10 FLAC 1001 with a Rice code, which is why this predictor does not encode the samples and provides no compression. Verbatim is essentially a zero-order predictor of the audio samples. The advantage of the verbatim predictor is fast decoding. If the decoder is told that a certain block or subblock employs this predictor, the decoder simply inputs the audio samples from the compressed stream without any decoding. Verbatim is the baseline against which the other predictors are measured. A good FLAC encoder should compress each subblock twice, first with the prediction method selected by the user, and then with the verbatim predictor. The smaller result should be written on the compressed stream. The verbatim predictor should do at least as well as any other predictor when the original audio is random or close to random. Constant. Sound is a wave. We hear sound when air vibrations strike our ear. Such vibrations are converted by a microphone to an electric wave (voltage that varies with time), and this wave is sampled to produce the audio samples. In the absence of sound, a microphone outputs a constant voltage (normally zero but possibly nonzero), which becomes a long sequence of identical audio samples; a digital silence. A sophisticated FLAC encoder should check each subblock for digital silence before it is encoded. When such a subblock is found, it is compressed by run-length encoding. Fixed linear predictor. This is a simple predictor that fits a polynomial to the audio samples (this method is selected if the FLAC user-controlled option -l is set to zero). Prediction by polynomial fit is faster than LPC (the next method) but usually results in files being 5-10% larger. The polynomial fit is similar to that employed by Shorten (zero-order to third-order prediction, Section 10.9), but also includes a fourth-order polynomial. The only side information included in the compressed stream is the order of prediction, a 3-bit integer with values of 0, 1, 2, 3, or 4. Fourth-order polynomial prediction is discussed in Section 10.10.1. FIR Linear prediction. This is a more sophisticated method that’s also referred to as general linear predictive coding (LPC) (see Section 7.30) and [Rabiner and Schafer 78]). This predictor is used if option -l is set to a value between 1 and 32. This value becomes the maximum order of the LPC. Generally, the larger the maximum order, the more accurate the prediction (fit) provided by LPC, but the slower the encoder. However, increasing the order (normally above 9) brings diminishing returns and may even degrade the LPC fit and thereby decrease compression. LPC is described in Section 10.10.2, and the following is quoted from the official FLAC site The reference encoder uses the Levinson-Durbin algorithm for calculating the LPC coefficients from the autocorrelation coefficients, and the coefficients are quantized before computing the residual. Whereas encoders such as Shorten use a fixed quantization for the entire input, FLAC allows the quantized coefficient precision to vary from subframe to subframe. The FLAC reference encoder estimates the optimal precision to use based on the block size and dynamic range of the original signal. Another difference between the last two methods is that the parameter for polynomial fit can be written on the compressed stream as a 3-bit number, whereas the size of the LPC parameters depends on the sample size (number of bits per audio sample) and the maximum LPC order. 1002 10. Audio Compression Residue encoding. Once the audio samples for a subblock have been computed, they are subtracted from the original samples to obtain the residues. It has been observed experimentally (see Section 10.2.1 for such an experiment) that the residues often have a Laplacian distribution (Equation (7.40) and Figure 7.143b), and it is known that the Rice codes (Section 10.9) are ideal for such a distribution. Thus, FLAC encodes the residues with Rice codes. Recall that each original block of audio samples becomes a frame on the compressed stream and each subblock becomes a subframe. Most of the content of a subframe are the Rice codes. The Rice codes depend on the choice of a single parameter m, and the best value for the parameter depends on the precise shape of the Laplace distribution. The FLAC encoder estimates m based on the statistical distribution of the residues, by m = log2(loge 2)E[|ri|] where E is the expected value of the residues ri (this estimate was originally proposed for Shorten in [Robinson 94]). Even more, the encoder can vary the Rice parameter inside a subframe if the distribution of the Rice codes varies significantly. The residues in a subframe can be broken up into partitions where each partition uses a different Rice parameter. The user-controlled option -r can be assigned values m and n (for min and max). The FLAC encoder then tries to divide each subframe into 2m partitions, then into 2m+1 partitions, and so on, up to 2n partitions. In each try, partitions are assigned different values of the Rice parameter, and the encoder measures the final size of the subframe. When all the sizes are known, the encoder selects the partitioning scheme that produces the smallest subframe. This process is slow and becomes even slower when the user selects values m and n that are very different. The FLAC literature recommends to set both m and n to 2, unless the block size is large, in which case the recommended values are m = n where n satisfies blocksize/(2n) = 128. Values m = 0 and n = 16 result in maximum optimization, but encoding is slow. The last encoding step is framing. The encoder prepares the frames and writes each on the compressed stream. Each frame contains a frame header, the Rice codes, and the frame footer. Format of FLAC Output. FLAC may be modified in the future, so the format of its output stream contains some empty, reserved spaces. In the future, these spaces may contain a FLAC version number and other information relating to as-yet nonexistent features. Certain fields are limited to only certain bit patterns, while other patterns are invalid. This contributes to the integrity of the output. All the numbers included in the output are integers and are in big-endian format. The output of FLAC starts with the marker fLaC, followed by one or more metadata blocks. The STREAMINFO metadata is mandatory and there may be up to 128 different metadata blocks. Currently, the following types of metadata are defined: The term Big Endian means that the high-order byte (the big end) of a number or a string is stored in memory at the lowest address (it comes first). For example, given the 4-byte number b3b2b1b0, if the most-significant byte b3 is stored at address A, then the least-significant byte b0 will be stored at address A + 3. STREAMINFO. This metadata block contains information about the entire output stream, such as the sampling rate, number of audio channels, and the total number of samples. STREAMINFO must be the first metadata block. 10.10 FLAC 1003 APPLICATION. A metadata block for use by third-party applications. The only mandatory field is a 32-bit identifier. Anyone can apply to the FLAC development team for such an ID. The remainder of an APPLICATION block is specified by the registered application. PADDING. This is a special type of metadata block that allows for an arbitrary amount of padding. The content of such a block is meaningless. This type of metadata is used in cases where the user plans to edit metadata after encoding. The user can instruct the encoder to reserve a PADDING block of sufficient size for later inclusion of metadata. SEEKTABLE. A metadata block with a table for storing seek points. There can be only one SEEKTABLE in an output stream, but the table can have any number of seek points. Seeking a particular point in a compressed audio file is especially useful. It is possible to skip to any given sample in a FLAC stream without a seek table, but the time it takes to skip is unpredictable because the bitrate may vary widely within a stream. Including seek points in the output stream significantly reduces the skip time. Each seek point occupies 18 bytes, so 1% resolution within a stream adds 100 Χ 18 = 1800 bytes to the output. VORBIS-COMMENT. This block implements the Vorbis comment specification. It contains a list of human-readable name/value pairs encoded in UTF-8. The VORBISCOMMENT metadata is the only officially-supported tagging mechanism in FLAC. There may be only one VORBIS-COMMENT block in an output stream. CUESHEET. This useful type of metadata block serves for data that can be used in a cue sheet. It supports track and index points that are compatible with the CD digital audio disc standard, as well as other CD-DA metadata. The CUESHEET block is especially useful for backing up CD-DA discs, but it seems that its normal use is as a general purpose cueing mechanism for audio playback. The metadata blocks are followed by the frames where each frame starts with a header. The frame header starts with a sync code, and contains the side information needed by the decoder, such as the sampling rate, number of bits per sample, number of audio channels, a frame/sample number, and an 8-bit CRC of the frame header itself. The 14-bit sync code 11111111111110 starts each frame. This code is important because the decoder may have to start decoding inside the compressed stream and must therefore have a way to locate the start of a frame. Unfortunately, the Rice codes inside a frame feature arbitrary bit patterns that may look like a sync code. Thus, when the decoder finds the bit pattern of a sync, it has to read the rest of the header and verify its CRC to make sure that this is a frame header. The sampling rate, number of bits per sample, and number of audio channels are the same for the entire compression job, but they appear in the header of every frame because the decoder may have to start decoding in midstream and may not have a chance to read the STREAMINFO metadata block located at the start of the output stream. The frame/sample number in the header is either the number of the first sample in the frame (in cases where frames have different sizes) or the number of the frame itself (in cases where frames have fixed sizes). The sync code, CRC, and frame/sample 1004 10. Audio Compression number allow resynchronization of the decoder in case of errors. They also allow seeking even if no seek points have been specified. Because the same parameters have to appear in every frame header, the header is merely overhead and it adversely affects the compression performance. To keep the headers small, at least in those cases where common parameters are used, FLAC employs lookup tables for the most-commonly-used values of certain parameters. For example, the sampling rate is stored in the frame header as a 4-bit code. Eight of the 16 possible codes (codes 4 through 11) specify the commonly-used sampling rates of 8, 16, 22.05, 24, 32, 44.1, 48, and 96 kHz. Three more codes direct the decoder to find the sampling rate at the end of the frame header. They are the following: Code 1100, find 8-bit sampling rate (in kHz) in end of header. Code 1101, find 16-bit sampling rate (in Hz) in end of header. Code 1110, find 16-bit sampling rate (in tens of Hz) in end of header. Code 1111 is invalid (to reduce the chance of sync fooling), and codes 0–3 are reserved. The same method is used to specify the block size and bits per sample. The (Rice encoded) subframes follow the header. Each subframe has its own header (with the attributes of the subframe, such as prediction method and order, and residual coding parameters), and there is one subframe for each audio channel. Notice that the subframes are not interleaved. The decoder therefore needs to reserve buffers large enough to read and store the subframes of a frame, and then loop over the buffers and decode a Rice code from each subframe in turn. The subframes are followed by a pad of zero bits, if necessary, to complete the last byte of the last subframe (recall that the Rice codes are variable-length, so the total length of a subframe may not be an integer multiple of eight bits). The frame footer contains a 16-bit CRC of the entire encoded frame, for robust error detection. If the reference decoder detects a CRC error, it generates a silent block. Other decoders may display an error message. 10.10.1 Fourth-Order Polynomial Prediction Section 10.9 develops the expressions for linear predictors of orders 0 through 3 (see also Figure 10.27). Extending these concepts to a 4th-order linear predictor is straightforward. We start with the four points P4 = (t-4, s4), P3 = (t-3, s3), P2 = (t-2, s2), and P1 = (t - 1, s1) and construct a degree-3 polynomial that passes through those points (the points are selected such that their x coordinates correspond to time and their y coordinates are audio samples). A natural choice is the nonuniform cubic Lagrange interpolation polynomial Q3nu(t) = 3 i=0 Pi+1L3i (t) whose coefficients are given by (see, for example, [Salomon 99] p. 204, Equations 4.17 and 4.18) L3i (t) = 73 j=i(t - tj) 73 j=i(ti - tj) , for 0 . i . 3. The Mathematica code of Figure 10.30 performs the computations and produces Q(t) = - 1 6 (t-1)(t-2)(t-3)P4+ 1 2t(t-2)(t-3)P3- 1 2t(t-1)(t-3)P2+ 1 6t(t-1)(t-2)P1. 10.10 FLAC 1005 It is easy to verify that Q(0) = P4, Q(1) = P3, Q(2) = P2, and Q(3) = P1. Extrapolating to t = 4 yields Q(4) = 4P1-6P2+4P3-P4, and when this is translated to audio samples the result is ˆs4(t) = 4s(t - 1) - 6s(t - 2) + 4s(t - 3) - s(t - 4) (10.7) [compare with Equation (10.5)]. When this prediction is subtracted from the current audio sample s(t), the residue is e4(t) = s(t) - ˆs4(t) = s(t) - 4s(t - 1) + 6s(t - 2) - 4s(t - 3) + s(t - 4). (10.8) This is a simple arithmetic expression that involves only four additions and subtractions. (* Uniform Cubic Lagrange polynomial for 4th-order prediction in FLAC *) Clear[Q,t]; t0=0; t1=1; t2=2; t3=3; Q[t_] := Plus @@ { ((t-t1)(t-t2)(t-t3))/((t0-t1)(t0-t2)(t0-t3))P4, ((t-t0)(t-t2)(t-t3))/((t1-t0)(t1-t2)(t1-t3))P3, ((t-t0)(t-t1)(t-t3))/((t2-t0)(t2-t1)(t2-t3))P2, ((t-t0)(t-t1)(t-t2))/((t3-t0)(t3-t1)(t3-t2))P1} Figure 10.30: Code for a Lagrange Polynomial.  Exercise 10.7: Check the performance of the 4th-order prediction developed here. Select four correlated items and compare the prediction of Equation (10.7) to the actual value of the next correlated item. 10.10.2 Linear Predictive Coding (LPC) Given a set of correlated values {ai}, linear predictive coding (LPC) is a sophisticated method to predict any set element ai from its n immediate predecessors ai-1 through ai-n, where n is a parameter. The idea is (see also Section 7.30) to try various sets of coefficients cj and select the set that minimizes the difference di =.. ai - n j=1 cjai-j.. 2 . (10.9) One reference for LPC is [Rabiner and Schafer 78]. First, some mathematical background. In statistics and probability we deal with random variables and define useful quantities such as average and variance. Given a random variable V that takes values vi, we denote the probability that V will have a specific value v by P(V = v). The average of V is (n i vi)/n, but statisticians also define a related quantity E called the expectation (or expected value) of V . The expectation E[V ] of random variable V is the sum of all the values vi of V , each multiplied by its probability. Thus, E[V ] = i viP(V = vi). 1006 10. Audio Compression The average and expectation of a random variable are often, but not always, the same. In many cases, the probability of a value equals its frequency of occurrence. When this is true, then the average and expectation are the same. Otherwise, they are different. The autocorrelation RV (d) of a random variable V is the correlation of V with a copy of itself, shifted d positions. For example, given an array a = (a1, a2, . . . , an) of n values, we construct the two arrays x = (a1, a2, . . . , an-1) and y = (a2, a3, . . . , an) of n - 1 values each, and compute the Pearson correlation coefficient R of x and y. This becomes the autocorrelation Ra(1) of array a with a shift of 1 position. Those unfamiliar with correlation may consult any text on probability and/or statistics. The short document [corr.pdf 02] may also be of help. What is important for the purpose of this discussion is that (under certain assumptions) the autocorrelation and expectation of a random variable are related by RV (k) = E[ViVi+k]. Now, back to linear predictive coding. In order to find the set of coefficients {cj} that minimizes Equation (10.9), we need to differentiate the expected value of di with respect to a coefficient cp, set the derivative to zero, and do this for all possible values of p, from 1 to n. The result is -2.. .. ai - n j=1 cjai-j.. ai-p.. = 0, for 1 . p . n, or n j=1 cjE[ai-jai-p] = E[aiai-p], for 1 . p . n. Replacing the expectations with autocorrelation coefficients yields the system of n linear equations ...... R(0) R(1) R(2) . . . R(n - 1) R(1) R(0) R(1) . . . R(n - 2) R(2) R(1) R(0) . . . R(n - 3) ... ... ... ... R(n - 1) R(n - 2) R(n - 3) . . . R(0) ...... ...... c1 c2 c3 ... cn ...... =...... R(1) R(2) R(3) ... R(n) ...... , with the n coefficients cj as the unknowns. This can be written compactly as RC = P and can easily be solved by Gaussian elimination or by inverting matrix R. The point is that matrix inversion requires in general O(n3) operations, but ours is not a general case, because our matrix R is special. It is easy to see that each diagonal of R consists of identical elements. Such a matrix is called a Toeplitz matrix, after its originator, Otto Toeplitz, and it can be inverted by a number of specialized, efficient algorithms. In addition, our R is also symmetric. FLAC employs a recursive, efficient method, known as the Levinson-Durbin algorithm, to solve this system of n equations. This algorithm was first proposed by Norman Levinson in 1947 [Levinson 47], improved by J. Durbin in 1960 [Durbin 60], and further improved by several researchers. In its present form it requires only 3n2 multiplications. A description of this algorithm is outside the scope of this book, but can be found, together with a derivation, in [Parsons 87]. 10.11 WavPack 1007 Otto Toeplitz (1881–1940) came from a Jewish family that produced several teachers of mathematics. Both his father, Emil Toeplitz, and his grandfather, Julius Toeplitz, taught mathematics in a Gymnasium and they also both published mathematics papers. Otto was brought up in Breslau and attended a Gymnasium in that city. His family background made it natural that he also should study mathematics. During his long, productive carrer, Toeplitz studied algebraic geometry, integral equations, and spectral theory. He worked also on summation processes, infinite-dimensional spaces and functions on them. In the 1930s he developed a general theory of infinite dimensional spaces and criticized Banach’s work as being too abstract. Toeplitz operators and Toeplitz matrices bear his name. We would like to thank Josh Coalson for his help with this section. 10.11WavPack (This section written by David Bryant, WavPack’s developer.) WavPack [WavPack 06] is a completely open, multiplatform audio compression algorithm and software that supports three compression modes, lossless, high-quality lossy, and a unique hybrid compression mode. It handles integer audio samples up to 32-bits wide and also 32-bit IEEE floating-point data [IEEE754 85]. The input stream is partitioned by WavPack into blocks that can be either mono or stereo and are generally 0.5 seconds long (but the length is actually flexible). Blocks may be combined in sequence by the encoder to handle multichannel audio streams. All audio sampling rates are supported by WavPack in all its modes. WavPack defaults to the lossless mode. In this mode, the audio samples are simply compressed at their full resolution and no information is discarded. WavPack generally provides a reasonable compromise between compression ratio and encoding/decoding speed. However, for specific applications, an alternate compromise may be more desirable. For this reason, WavPack incorporates an optional “fast” mode that is fast and entails only a small penalty in compression performance and a “high” mode that aims for maximum compression (at a somewhat slower pace). The lossy mode employs no subband coding or psychoacoustic noise masking. Instead, it is based on variable quantization in the time domain combined with mild noise shaping. This mode can operate from bitrates as low as 2.22 bits per sample up to fully lossless and offers considerably more flexibility and performance than the similar, but much simpler, ADPCM (Section 10.6). This makes the lossy mode ideal for high-quality audio encoding where the data storage or bandwidth requirements of the lossless mode might be prohibitive. 1008 10. Audio Compression Finally, the hybrid mode combines the lossless and lossy modes into a single operation. Instead of a single output file being generated from a source, two files are generated. The first, with the extension wv, is a smaller lossy file that can be played on its own. The second is a “correction” file (wvc) that, when combined with the first, provides lossless restoration of the original audio. The two files together have essentially the same size as the output produced by the pure lossless mode, so the hybrid mode generates very little additional overhead. Efficient hardware decoding was among the design goals of WavPack, and this weighed heavily (along with patent considerations) in the design decisions. As a result, even the most demanding WavPack mode will easily decode CD quality audio in real time on an ARM7 processor running at 75 MHz (and the default modes use considerably less than that). audio samples left right joint stereo processing mid side residuals feedback (lossy + hybrid) (hybrid) coded residual ‘correction’ codewords interleaved codewords entropy coder multipass decorrelation mid side Figure 10.31: The WavPack Encoder. Figure 10.31 is a block diagram of the WavPack encoder. Its main parts are the joint-stereo processing (removing inter-channel correlations), multipass decorrelation (removing intra-channel correlations between neighboring audio samples), and the entropy coder. Decorrelation The decorrelation passes are virtually identical in all three modes of WavPack. The first step in the decorrelation process is to convert the left and right stereo channels to the standard difference and average (also referred to as side and mid) channels by s(k) = l(k)-r(k) and s(k) = int(l(k)+r(k))/2. Notice that the integer division by 2 discards the least-significant bit of the sum l(k)+r(k), but this bit can be reconstructed from the difference, because a sum A + B and a difference A - B have the same leastsignificant bit. The second decorrelation step is prediction. WavPack performs multiple passes of very simple, single-tap linear predictive filters that employ the sign-sign LMS method for adaptation. (The sign-sign LMS method [adaptiv9 06] is a variation on the standard least-mean-squares method.) Having only one adjustable weighing factor per filter eliminates problems of instability and non-convergence that can occur when multiple taps are updated from a single result. The overall computational complexity of the 10.11 WavPack 1009 process is simply controlled by the total number of filtering passes made. The default case performs five passes, as this provides a good compromise of compression vs. speed. In the “fast” mode, only two passes are made, while the “high” mode incorporates the maximum allowed 16 passes. There are actually 13 variations of the filter depending on what sample value (or linear function of two sample values) is used as the input value u(k) to the prediction filter. These variations are identified by a parameter named “term.” In the case of a single channel, filtering depends on the value of this parameter as follows u(k) =... 1 . term . 8, s(k - term), term = 17, 2s(k - 1) - s(k - 2), term = 18, 3s(k - 1) - s(k - 2)/2.0, where s(k) is the sample to be predicted and the recent sample history is s(k - 1), s(k - 2),. . . , s(k - 8). For the dual channel case (which may be left and right or mid and side) there are three more filter variations available that use cross-channel information if(term = -1), u(k) = s(k), u(k) = s(k - 1), if(term = -2), u(k) = s(k - 1), u(k) = s(k), if(term = -3), u(k) = s(k - 1), u(k) = s(k - 1). Once u(k) has been determined, an error (or residual) value e(k) is computed as the difference e(k) = s(k)-w(k)u(k), where w(k) is the current filter weight value. The weight generally varies between +1 for highly correlated samples and 0 for completely uncorrelated samples. It can also become negative if there is a negative correlation between the sample values. Negative correlation would exist if (-1Χu(k)) was better than u(k) as a predictor of s(k). This often happens with samples with large amounts of high frequencies. Finally, the weight is updated for the next sample based on the signs of the filter input and the residual w(k + 1) = w(k) + d · sgn(u(k)) · sgn(e(k)), where parameter d is the step-size of the adaptation, a small positive integer. (The sgn function used by WavPack returns +1, 0, or -1.) The rule for adapting w(k) implies that if either u(k) or e(k) is zero, the weight is not updated. In cases where the input to the filter comes from the same channel as the sample being predicted (i.e., where “term” is positive), the magnitude of the weight is self limiting and is therefore allowed to exceed ±1. However, in the cross-channel cases (where “term” is negative), the weight could grow indefinitely, and so is clipped to ±1 after each adaptation. Here’s an example of an adaptation step. Assuming that the input u(k) to the predictor is 105, sample s(k) to be predicted is 100, the current filter weight w(k) is 0.8, and the current stepsize d is 0.01, we first compute the residual e(k) = s(k)-w(k)·u(k) = 100 - 105 · 0.8 = 16 and then update the weight w(k + 1) = w(k) + s · sgn(e(k)) · sgn(u(k)) = 0.8 + 0.01 · sgn(100) · sgn(16) = 0.8 + 0.01 · 1 · 1 = 0.81. 1010 10. Audio Compression We can see that if the original weight had been more positive (greater than 0.8), the prediction would have been closer to s(k) and the residual would have been smaller. Therefore, we increase the weight in this iteration. The magnitude of the stepsize is a compromise between fast adaptation and noisy operation around the optimum value. In other words, if the stepsize is large, the filter will adjust quickly to sudden changes in the signal, but the weight will jump around a lot when it is near the correct value. Too small a stepsize will cause the filter to adjust too slowly to rapidly changing audio. In the lossless mode, the results of the decorrelation (the residuals e(k)) are simply passed to the entropy coder for exact translation. In the lossy and hybrid modes, in contrast, the decorrelated values may be coded inexactly based on a defined maximum allowable error. In this situation, the actual value coded is returned from the entropy coder and this value is used to update the filter weight and generate the reconstructed sample value. This is required because during subsequent decoding the exact sample values will not be available, and we must operate the encoder with only the information that we will have at that time. Implementation All sample data is presented to the decorrelator as 32-bit signed integers, even if the source is smaller. This is done to allow common routines, but is also required from an arithmetic standpoint because, for example, 16-bit data might clip at some steps if restricted to 16-bit storage. The routines are designed in such a way that they will not overflow on sample data with up to 25 bits of significance. This allows for standard 24-bit audio and the 25-bit significand of 32-bit IEEE data. Any significance beyond this must be handled by side channel data. The filter weights are represented in 16-bit signed integers with 10 bits of fraction (for example 1024 = 0000010 . . . 00 3 45 6 10 2 represents 1.0), and the adjusting step-size may range from 1 to 7 units of that resolution. For input values that fit in 16-bit signed integers, it is possible to perform the rounded weight multiply with a single 16-bit Χ 16-bit = 32-bit operation by prediction = (sample * weight + 512) >> 10; Since “sample” and “weight” both fit in 16-bit integers, the signed product will fit in 32 bits. We add the 512 for rounding and then shift the result to the right 10 places to compensate for the 10 bits of fraction in the weight representation. This operation is very efficiently implemented in modern processors. For those cases where the sample does not fit in 16 bits, we need more than 32 bits in the intermediate result, and several alternative methods are provided in the source code. If a multiply operation is provided with 40 bits of product resolution (as in the ColdFire EMAC or several Texas Instruments DSP chips), then the above operation will work (although some scaling of the inputs may be required). If only 32-bit results are provided, it is possible to perform separate operations for the lower and upper 16 bits of the input sample and then combine the results with shifts and an addition. This has been efficiently implemented with the MMX extensions to the x86 instruction set. Finally, some processors provide hardware-assisted double-precision floating-point arithmetic which can easily perform the operation with the required precision. 10.11 WavPack 1011 As mentioned previously, multiple passes of the decorrelation filter are used. For the standard operating modes (fast, default, and high) stepsize is always 2 (which represents about 0.002) and the following terms are employed, in the order shown: fast mode terms = {17, 17}, default mode terms = {18, 18, 2, 3,-2}, high mode terms = {18, 18, 2, 3,-2, 18, 2, 4, 7, 5, 3, 6, 8,-1, 18, 2}. For single channel operation, the negative term passes (i.e. passes with cross-channel correlation) are skipped. Note, however, that the decoder is not fixed to any of these configurations because the number of decorrelation passes, as well as other information such as what terms and stepsize to use in each pass, is sent to the decoder as side information at the beginning of each block. For lossless processing, it is possible to perform each pass of the decorrelation in a separate loop acting on an array of samples. This allows the process to be implemented in small, highly optimized routines. Unfortunately, this technique is not possible in the lossy mode, because the results of the inexact entropy coding must be accounted for at each sample. However, on the decoding side, this technique may be used for both lossless and lossy modes (because they are essentially identical from the decorrelator’s viewpoint). Generally, the operations are not parallelizable because the output from one pass is the input to the next pass. However, in the two-channel mode the two channels may be performed in parallel for the positive term values (and again this has been implemented in MMX instructions). Entropy Coder The decorrelation step generates residuals e(k) that are signed numbers (generally small). Recall that decorrelation is a multistep process, where the differences e(k) computed by a step become the input of the next step. Because of this feature of WavPack, we refer to the final outputs e(k) of the decorrelation process as residuals. The residuals are compressed by the entropy encoder. The sequence of residuals is best characterized as a two-sided geometric distribution (TSGD) centered at zero (see Figure 3.49 for the standard, one-sided geometric distribution). The Golomb codes (Section 3.24) provide a simple method for optimally encoding similar one-sided geometric distributions. This code depends on a single positive integer parameter m. To Golomb encode any nonnegative integer n, we first divide n into a magnitude mag = int(n/m) and a remainder rem = n mod m, then we code the magnitude as a unary prefix (i.e. mag 1’s followed by a single 0) and follow that with an adjusted-binary code representation of the remainder. If m is an integer power of 2 (m = 2k), then the remainder is simply encoded as the binary value of rem using k bits. If m is not an integer power of 2, then the remainder is coded in either k or k + 1 bits where 2k < m < 2k+1. Because the values at the low side of the range are more probable than values at the high side, the shorter codewords are reserved for the smaller values. Table 10.32 lists the adjusted binary codes for m values 6 through 11. The Rice codes are a common simplification of this method, where m is a power of 2. This eliminates the need for the adjusted binary codes and eliminates the division 1012 10. Audio Compression remainder when when when when when when to code m = 6 m = 7 m = 8 m = 9 m = 10 m = 11 0 00 00 000 000 000 000 1 01 010 001 001 001 001 2 100 011 010 010 010 010 3 101 100 011 011 011 011 4 110 101 100 100 100 100 5 111 110 101 101 101 1010 6 111 110 110 1100 1011 7 111 1110 1101 1100 8 1111 1110 1101 9 1111 1110 10 1111 Table 10.32: Adjusted Binary Codes for Six m Values. required by the general Golomb codes (because division by a power of 2 is implemented as a shift). Rice codes are commonly used by lossless audio compression algorithms to encode prediction errors. However, Rice codes are less efficient when the optimum m is midway between two consecutive powers of 2. They are also less suitable for lossy encoding because of discontinuities resulting from large jumps in the value of m. For these reasons Rice codes were not chosen. Before a residual e(k) is encoded, it is converted to a nonnegative value by means of if e(k) < 0, e(k) = -e(k)+1. The original sign bit of e(k) is appended to the end of the coded value. Note that the sign bit is always written on the compressed stream, even when the value to be encoded is zero, because -1 is also coded as zero. This is less efficient for very small encoded values where the probability of the value zero is significantly higher than neighboring values. However, we know from long experience that an audio file may have long runs of consecutive zero audio samples (silence), but relatively few isolated zero audio samples. Thus, our chosen method is slightly more efficient when zero is not significantly more common than its neighbors. We show later that WavPack encodes runs of zero residuals with an Elias code. The simplest method for determining m is to divide the residuals into blocks and determine the value of m that encodes the block in the smallest number of bits. Compression algorithms that employ this method have to send the value of m to the decoder as side information. It is included in the compressed stream before the encoded samples. In keeping with the overall philosophy of WavPack, we instead implement a method that dynamically adjusts m at each audio sample. The discussion on page 164 (and especially Equation (3.12)) shows that the best value for m is the median of the recent sample values. We therefore attempt to adapt m directly from the current residual e(k) using the simple expression if e(k) . m(k) then m(k + 1) = m(k) + int m(k) + 127 128 , 10.11 WavPack 1013 else m(k + 1) = m(k) - int m(k) + 126 128 . (10.10) The idea is that the current median m(k) will be incremented by a small amount when a residual e(k) occurs above it and will be decremented by an almost identical amount when the residual occurs below it. The different offsets (126 and 127) for the two cases were selected so that the median value can move up from 1, but not go below 1. This works very well for large values, but because the median is simply an integer, it tends to be “jumpy” at small values because it must move at least one unit in each step. For this reason the actual implementation adds four bits of fraction to the median that are simply ignored when the value is used in a comparison. This way, it can move smoothly at smaller values. It is interesting to note that the initial value of the median is not crucial. Regardless of its initial value, m will move towards the median of the values it is presented with. In WavPack, m is initialized at the start of each block to the value that it left off on the previous block, but even if it started at 1 it would still work (but would take a while to catch up with the median). There is a problem with using an adaptive median value in this manner. Sometimes, a residual will come along that is so big compared to the median that a Golomb code with a huge number of 1’s will be generated. For example, if m was equal to 1 and a residual of 24,000 occurred, then the Golomb code for the sample would result in mag = int(24,000/1) = 24,000 and rem = 24,000 mod 1 = 0 and would therefore start with 24,000 1’s, the equivalent of 3,000 bytes of all 1’s. To prevent such a situation and avoid ridiculously long Golomb codes, WavPack generates only Golomb codes with 15 or fewer consecutive 1’s. If a Golomb code for a residual e(k) requires k consecutive 1’s, where k is greater than or equal 16, then WavPack encodes e(k) by generating 16 1’s, followed by the Elias gamma code of k (Code C1 of Table 3.20). When the residuals are distributed geometrically, this situation is exceedingly rare, so it does not affect efficiency, but it does prevent extraordinary and rarely-occurring cases from drastically reducing compression. Recall that each encoded value is followed by the sign bit of the original residual e(k). Thus, an audio sample of zero is inefficiently encoded as the two bits 00. Because of this, WavPack employs a special method to encode runs of consecutive zero residuals. A run of k zeros is encoded in the same Elias gamma code preceded by a single 1 to distinguish it from a regular code (this is done only when the medians are very small to avoid wasting a bit for every sample.) One source of less than optimum efficiency with the method described so far is that sometimes the data is not distributed in an exactly geometrical manner. The decay for higher values may be faster or slower than exponential, either because the decorrelator is not perfect or the audio data has uncommon distribution characteristics. In these cases, it may be that the dynamically estimated median is not a good value for m, or it may be that no value of m can produce optimal codes. For these cases, a further refinement has been added that I have named Recursive Golomb Coding. In this method, we calculate the first median as in Equation (10.10) and denote it by m. However, for values above the median we subtract m from the value and calculate a new median, called m, that represents the 3/4 population point of samples. Finally, 1014 10. Audio Compression we calculate a third median (m) for the sample values that lie above the second median (m), and we use this for all remaining partitions (Figure 10.33). The first three coding zones still represent 1/2, 1/4, and 1/8 of the residuals, but they don’t necessarily span the same number of possible values as they would with a single m. However, this is not an issue for the decoder because it knows all the medians and therefore all the span sizes. distribution wider distribution geometric residuals 1/4 residuals 1/2 residuals 1/8 residuals 1/16 m m’ m’’ m’’ 0(ab)s 10(ab)s 110(ab)s 1110(ab)s ... Figure 10.33: Median Zones for Recursive Golomb Coding. In practice, recursive Golomb coding works like this. If the residual being coded is in the first zone, fine. If not, then subtract m from the residual and pretend we’re starting all over. In other words, first we divide the residuals into two equal groups (under m and over m). Then, we subtract m from the “over m” group (so they start at zero again) and divide those into two groups. Then, we do this once more. This way, the first several zones are guaranteed to have the proper distribution because we calculate the second and third medians directly instead of assuming that they’re equal in width to the first one. Table 10.34 lists the first few zones where (ab) represents the adjusted-binary encoding of the remainder (residual modulo m) and S represents the sign bit. range prob. coding 0 . residualm(k) then m(k + 1) = m(k) + 5 · int m(k) + 127 128 , else m(k + 1) = m(k) - 2 · int m(k) + 125 128 . The idea is that the median is bumped up five units every time the value is higher, and bumped down two units when it is lower. If there are two ups for every five downs, the median stays stationary. Because the median moves down a minimum of two units, we consider 1 or 2 to be the minimum (it will not go below this). Now the first region contains 5/7 of the residuals, the second has 10/49, the third 20/343, and so on. Using these enlarged median zones, the minimum bitrate drops from 1016 10. Audio Compression 3 to 2.4 bits per residual, as shown by the following infinite sum 1 + . n=1 5nΧ2n-1 7n . There is an efficiency loss with these unary codes because the number of 1’s and 0’s in the datastream is no longer equal. There is a single 0 for each residual, but since there are only 1.4 total bits per residual (not counting the sign bit), there must be only an average of 0.4 1’s per residual. In other words, the probability for 0’s is 5/7 and the probability for 1’s is 2/7. Using the log2 method to determine the ideal number of bits required to represent those symbols, we obtain the numbers listed in Table 10.36a and derive an optimum multibit encoding scheme (Table 10.36b). symbol freq log2 (1/freq) 0 5/7 0.485427 1 2/7 1.807355 input ideal output sequence freq # bits sequence 00 25/49 0.970854 0 01 10/49 2.292782 10 1 14/49 1.807355 11 (a) (b) Table 10.36: Optimum Translation of Unary Magnitudes. input input net* output output net* sequence freq bits input sequence bits output 00 25/49 2 50/49 0 1 25/49 01 10/49 2 20/49 10 2 20/49 1 14/49 1 14/49 11 2 28/49 totals 1 84/49 73/49 *The “net” values are the frequency multiplied by the number of bits. Table 10.37: Net Savings from Unary Translation. Table 10.37 shows that the average number of unary magnitude bits transmitted is reduced by the translation to (73/84)Χ1.4 = 1.21666 . 1.22 bits per sample. We use the 1.4 figure because only the unary magnitude data is subject to this translation; adding back the untranslated sign bit gets us to about 2.22 bits per sample. This represents the encoder running “flat out” lossy (i.e., no remainder data sent) however, the method is so efficient that it is used for all modes. In the lossless case, the complete adjustedbinary representation of the remainder is inserted between the magnitude data and the terminating sign bit. 10.12 Monkey’s Audio 1017 In practice, we don’t normally want the encoder to be running at its absolute minimum lossy bitrate. Instead, we would generally like to be somewhere in-between full lossy mode and lossless mode so that we send some (but not all) of the remainder value. Specifically, we want to have a maximum allowed error that can be adjusted during encoding to give us a specific signal-to-noise ratio or constant bitrate. Sending just enough data to reach a specified error limit for each sample (and no more) is optimal for this kind of quantization because the perceived noise is equivalent to the RMS level of the error values. To accomplish this, we have the encoder calculate the range of possible sample values that satisfy the magnitude information that was just sent. If this range is smaller than the desired error limit, then we are finished and no additional data is sent for that sample. Otherwise, successive bits are sent to the datastream, each one narrowing the possible range by half (i.e., 0 = lower half and 1 = upper half) until the size of the range becomes less than the error limit. This binary process is done in lockstep by the decoder, to remain synchronized with the bitstream. At this point it should be clear how the hybrid lossless mode splits the bitstream into a lossy part and a correction part. Once the specified error limit has been met, the numeric position of the sample value in the remaining range is encoded using the adjusted binary code described above, but this data goes into the “correction” stream rather than the main bitstream. Again, on decoding, the process is exactly reversed so the decoder can stay synchronized with the two bitstreams. With respect to the entropy coder, the overhead associated with having two files instead of one is negligible. The only cost is the difference resulting from coding the lossy portion of the remainder with the binary splitting method instead of the adjusted binary method. The rest of the cost comes on the decorrelator side where it must deal with a more noisy signal. More information on the theory behind and the implementation of WavPack can be found in [WavPack 06]. 10.12 Monkey’s Audio Monkey’s audio is a fast, efficient, free, lossless audio compression algorithm and implementation (for the Windows operating system) by Matt Ashland [monkeyaudio 06]. This method also has the following important features: Error detection. The implementation incorporates CRCs in the compressed stream to detect virtually all errors. Tagging support. The algorithm inserts tags in the compressed stream. This makes it easy for a user to manage and catalog an entire collection of monkey-compressed audio files. External support. Monkey’s audio can also be used as a front-end to a commandline encoder. The current GUI tool allows different command line encoders to be used by creating a simple XML script that describes how to pass parameters and what return values to look for. 1018 10. Audio Compression The source code is freely available. Software developers can use it (modified or not) in other programs without restrictions. The first step in monkey’s audio is to transform the left and right stereo channels to X (mid or average) and Y (side or difference) channels according to X = (L+R)/2 and Y = L- R. The inter correlation between L and R implies that the mid channel X is similar to both L and R, while the side channel Y consists of small numbers. Notice that the elements (transform coefficients) of X and Y are integers but are no longer audio samples. Also, the transform is reversible even though the division by 2 causes the loss of the least-significant bit (LSB) of X. This bit can be reconstructed by the decoder because it is identical to the LSB of Y. (Notice that the LSBs of any sum A + B and difference A - B are the same.) The next step is prediction. This step exploits the intra correlation between consecutive elements in the mid and side channels. Monkey’s audio employs a novel method for linear prediction that consists of a fixed first-order predictor followed by multiple adaptive offset filters. There are varying degrees of neural net filtering, depending on the compression level—up to three consecutive neural-net filters with 1024 sample windows at the highest compression. In its last step, monkey’s audio encodes the residuals with a range-style coder (Section 5.10.1). The compressed stream is organized in frames, where each frame has a 31-bit CRC (a standard 32-bit CRC with the most-significant bit truncated). The encoder also computes an MD5 checksum for the entire file. The decoder starts by verifying this checksum. A failed verification indicates an error, and the decoder locates the error by verifying the CRCs of the individual frames. The best lossless audio compressor around. . . written by the best monkey around. Great job, Matt! —From http://www.ashlands.net/Links.htm. 10.13 MPEG-4 Audio Lossless Coding (ALS) MPEG-4 Audio Lossless Coding (ALS) is the latest addition to the family of MPEG-4 audio codecs. ALS can input floating-point audio samples and is based on a combination of linear prediction (both short-term and long-term), multichannel coding, and efficient encoding of audio residues by means of Rice codes and block codes (the latter are also known as block Gilbert-Moore codes, or BGMC [Gilbert and Moore 59] and [Reznik 04]). Because of this organization, ALS is not restricted to the encoding of audio signals, and can efficiently and losslessly compress other types of one-dimensional, fixed-size, correlated signals, such as medical (ECG-EKG and EEG) and seismic data. Because of its lossless nature and high compression gain, its developers envision the following audio compression applications for ALS: Archival (audio archives in broadcasting, sound studios, record labels, and libraries) Studio operations (storage, collaborative working, and digital transfer) 10.13 MPEG-4 Audio Lossless Coding (ALS) 1019 High-resolution disc formats (CD, DVD, and future formats) Internet distribution of audio files Online music stores (downloading purchased music) Portable music players (an especially popular application) Currently, not all these applications employ ALS, because many believe that audio compression should always be lossy and there is no point in preserving every bit of audio, because the ear cannot distinguish between reconstructed lossy and reconstructed lossless sounds. There is an ongoing debate about the advantage of lossless audio compression over lossy methods and whether lossless audio compression is really necessary. An Internet search returns many strong opinions, ranging from “Not another losslessphile. When will you people realize that there is no point in lossless audio compression?” to “I spent all of last summer converting my 200 Gb audio collection to a lossy format, and now I have to redo it in lossless because of the low quality of mp3.” Here are two more opinions: “While trying to encode lossless some of it I came up with enormous files. For example with only 50 minutes of music, the Bach Inventions and Sinfonias with Kenneth Gilbert in Archiv ended up around 350 Mb with Ape. Apparently the sound of cembalo [harpsichord] is so rich in harmonics that it creates very complex sound waves (with a very high entropy indicator—my apologies, I’m translating a Greek term). That’s why many cembalo-solo mp3 files sound not-so-pleasant in mp3 (while other instruments with a more ‘rounded’ sound like the flute, sound just fine).” “For me actually harpsichord mp3 sounds better than other instruments simply because it is rich in harmonics, decoding to mp3 cuts off a lot of high frequencies, harpsichord has a lot so it still sounds good, other instruments, for example piano, sound much worse, human voice is very vulnerable to cutting too many high frequencies as well. Besides, you have to know how to make a good mp3. Some 256K/s bitrate mp3s are horrible while sometimes a 160K/s bitrate mp3 sounds really good. Of course, such low bitrates are almost always too low for orchestral music, in this case only lossless compression is enough (if you have a decent CD player and good ripping software of course).” The main reference for MPEG-4 ALS is [mpeg-4.als 06], a Web site maintained by Tilman Liebchen, a developer of this algorithm and the editor of the MPEG-4 ALS standard. MPEG-4 ALS (in this section we will refer to it simply as ALS) supports any audio sampling rates, audio samples of up to 32 bits (including samples in 32-bit floatingpoint format), and up to 216 audio channels. Tests performed by the ALS development team are described in [Liebchen et al. 05] and indicate that ALS, even in its simplest modes, outperforms FLAC. First, a bit of history. In July 2002, the MPEG committee issued a call for proposals for a lossless audio coding algorithm. By December 2002, seven organizations have submitted codecs that met the basic requirements. These were evaluated in early 2003 in terms of compression efficiency, complexity, and flexibility. In March 2003, the codec proposed by researchers from the Technical University of Berlin was selected as the first working draft. The following two years saw further development, followed by implementation and testing. This work culminated in the technical specifications of ALS, which 1020 10. Audio Compression were finalized in July 2005, and adopted by the MPEG committee and later by the ISO. The formal designation of ALS is the ISO/IEC 14496-3:2005/Amd 2:2006 international standard. The organization of the ALS encoder and decoder is shown in Figure 10.38. The original audio data is divided into frames, each frame is divided into audio channels, and each channel may be divided into blocks. The encoder may modify block sizes if this improves compression. Audio channels may be joined by either subtracting pairs of adjacent channels or by multi-channel coding, in order to reduce inter-channel redundancy. The audio samples in each block are predicted by a combination of short-term and long-term predictors to generate residues. The residues are encoded by Rice codes and block Gilbert-Moore codes that are written on the output stream. This stream also includes tags that allow random access of the audio at virtually any point. Another useful optional feature is a CRC checksum for better integrity of the output. Multiplexing Demultiplexing Encoder Decoder Short term prediction Short term prediction Long term prediction Long term prediction Joint channel coding Joint channel decoding Entropy coding Entropy decoding Control Data Compressed bitstream Frame/block partition Input Output Frame/block assembly fl Figure 10.38: Organization of the ALS Encoder and Decoder. Prediction. Linear prediction is commonly used in lossless audio encoders and its principles have been mentioned elsewhere in this book (see Sections 7.30, 10.9, and 10.10.1). A finite-impulse-response (FIR) linear predictor of order K predicts the current audio sample x(n) from its K immediate predecessors by computing the linear sum ˆx(n) = K k=1 hkx(n - k), where hk are coefficients to be determined. The residue e(n) is computed as the difference x(n) - ˆx(n) and is later encoded. If the prediction is done properly, the residues are decorrelated and are also small numbers, ideal for later encoding with a variable-length code. 10.13 MPEG-4 Audio Lossless Coding (ALS) 1021 The decoder obtains e(n) by reading its code and decoding it. It then computes ˆx(n) from the coefficients hk. Thus, the decoder needs to know these coefficients and this can be achieved in three ways as follows: 1. The coefficients are always the same. This is termed nth-order prediction and is discussed in Section 10.9 for n = 0, 1, 2, and 3 and in Section 10.10.1 for n = 4. 2. The coefficients are computed based on the K audio samples preceding x(n) (this is referred to as forward prediction) and may also depend on some residues preceding e(n) (backward prediction). These samples and residues are known to the decoder and it can compute the coefficients in lockstep with the encoder. 3. The coefficients are computed by the encoder based on all the audio samples in the current block. When the decoder gets to decoding audio sample x(n), it has only the samples preceding it and not those following x(n). Therefore, the coefficients have to be transmitted in the bitstream, as side information, for the use of the decoder. Naturally, they have to be efficiently encoded. Buffer Multiplexing Predictor Encoder Parcor Parcor Values Q coding Entropy coding Entropy to LPC Code indices Bitstream parcor values Quantized Estimate Residues Original Figure 10.39: Forward Adaptive Prediction in ALS. ALS adopts forward-adaptive prediction as in 3 above with a prediction-order K of up to 1023. Figure 10.39 is a block diagram of the prediction steps of the ALS encoder. An entire block of audio samples is read into a buffer, and the encoder executes the Levinson-Durbin algorithm (as also does FLAC, Section 10.10). This algorithm (see, for example, [Parsons 87] for a derivation) is recursive, where each iteration increments the order of the filter. The final order K becomes known in the last iteration. The algorithm produces a set of filter coefficients hi that minimizes the variance of the residues for the block. The Levinson-Durbin algorithm is adaptive because it computes a different set of filter coefficients and a different order K for each block, depending on the varying statistics of the audio samples and on the block size. The result is a set of filter coefficients that not only produce the smallest residues, but are themselves small, easy to compress, and therefore minimize the amount of side information in the bitstream. [The value of the predictor order K is important. A large K reduces the variance of the prediction error and results in a smaller bitrate Re for the encoded residues. On the 1022 10. Audio Compression other hand, a large prediction order implies more filter coefficients and consequently a higher bitrate Rc of the encoded coefficients (the side information). Thus, an important aspect of the Levinson-Durbin algorithm is to determine the value K that will minimize the total bitrate R(K) = Re(K) + Rc(K). The algorithm estimates both Re(K) and Rc(K) in each iteration and stops when their sum no longer decreases.] While computing the filter coefficients hi in an iteration, the Levinson-Durbin algorithm also computes a set of constants ki that are referred to as partial correlation (parcor) coefficients (also known as reflection coefficients, see [Rabiner and Schafer 78]). These coefficients have an interesting property. As long as their magnitudes are less than 1, they produce a stable filter. Therefore, it is preferable to quantize the parcor coefficients ki and use the quantized parcors to compute the desired coefficients hi. The quantized parcors are compressed and multiplexed into the bitstream to be used later by the decoder. The parcors are now used to compute a set of filter coefficients hi that are in turn used to predict all the audio samples in the block. The predicted estimates are subtracted from the actual samples, and the resulting residues are encoded and are also multiplexed into the bitstream. The ALS decoder (Figure 10.40) is much simpler because it doesn’t have to compute any coefficients. For each block, it demultiplexes the encoded parcors from the bitstream, decodes them, and converts them to filter coefficients hi that are then used to compute the estimates ˆx(n) for the entire block. Each encoded residue is then read from the bitstream, decoded, and converted to an audio sample. The complexity of computations in the decoder depends mostly on the order K of the prediction and this order varies from block to block. Demultiplexing Predictor Decoder Parcor coding Entropy to LPC Code indices Bitstream reconstruction Lossless Estimate Entropy Residues decoder Parcor values Figure 10.40: Decoding of Predictions in ALS. Quantization. ALS quantizes the parcors ri, not the filter coefficients hi, because it is known (see, for example, [Kleijn and Paliwal 95]) that the latter are very sensitive to even small quantization errors. In order to end up with a stable filter, the parcors are quantized to the interval [-1, 1] and the quantization method takes into account two important facts: (1) The parcors are less sensitive to quantization errors than the filter coefficients, but parcors close to +1 or -1 are sensitive. (2) The first two parcors r1 and r2, are normally close to -1 and +1, respectively, while the remaining parcors are normally smaller. Thus, the main quantization problem is with regard to the first two 10.13 MPEG-4 Audio Lossless Coding (ALS) 1023 parcors and is solved by the simple companding expression C(r) = -1 +2(r + 1). Figure 10.41 shows the behavior of this compander. Function C(r) is close to vertical at r = -1, implying that it is very sensitive to small values of r around -1. Thus, C(r) is an ideal compander for r1. Its complement, -C(-r), is close to vertical at r = +1 and is therefore an ideal candidate to compand r2. In order to simplify the computations, +C(-r2) is actually used to compand r2, leading to a companded value that is around -1. After companding, both parcors are quantized to seven bit values (a sign bit and six bits of magnitude) a1 and a2 using a simple, uniform quantizer. The final results are a1 = #64 -1 +2(r1 + 1)$, a2 = #64 -1 +2(-r2 + 1)$. The remaining parcors ri (i > 2) are not companded, but are quantized to seven bits by the uniform quantizer ai = 64ri. The 7-bit quantized parcors ai are signed, which places them in the interval [-64, +63]. This interval is centered at zero, but the average of the ais may be nonzero. To shrink them even further, the ais are re-centered around their most probable values and are then encoded with Rice codes. The result is that each quantized recentered parcor can be encoded with about four bits, without any noticeable degradation of the reconstructed audio. The parcor to LPC conversion shown in Figures 10.39 and 10.40 uses integer computations and can therefore be reconstructed to full precision by both encoder and decoder. -C(-r) C(r) -1 -0.5 -1 0.5 0.5 1 1 0 0 -0.5 r Figure 10.41: Compander Functions C(r) and -C(-r). Block Size Switching. As mentioned earlier, the input is divided into fixed-size frames. The frame size can be selected by the user depending on the audio sampling rate. For example, if the user prefers 43 ms of audio in a frame, then the frame size should be set to 2,048 samples if the sampling rate is 48 kHz, or 4,096 samples if the sampling rate is 96 kHz. Initially, each frame is one block, but a sophisticated encoder may subdivide a frame into blocks in order to adapt the linear prediction to variations 1024 10. Audio Compression in the input audio. This is referred to as block switching. In general, long blocks with high predictor orders are preferable for stationary segments of audio, while short blocks with lower orders are suitable for transient audio segments. A frame of size N can be subdivided into blocks of sizes N/2, N/4, N/8, N/16, and N/32, and the only restriction is that in each subdivision step a block of size n should be divided into two blocks of size n/2 each. Figure 10.42 shows some examples of valid (a–d) and invalid (e, f) block subdivisions. N N/2 N/2 N/2 N/4 N/4 N/4 N/4 N/4 N/4 N/4 N/4 N/4 N/4 N/4 N/4 N/8 N/8 N/8 N/8 (a) (b) (d) (e) (f) (c) Figure 10.42: Block Subdivision Examples. The ALS standard does not specify how an encoder is to determine block switching. Methods may range from no block switching, to switching by estimating audio signal characteristics, to complex algorithms that try many block decompositions and select the best one. The block subdivision (BS) scheme selected by the endocder has to be transmitted to the decoder as side information, but it does not increase the work of the decoder since the decoder still has to process the same number of audio samples per frame regardless of how the frame has been subdivided. Thus, block switching can be employed by a sophisticated encoder to significantly increase compression, without degrading decoder performance. Because of the rule governing block switching, the BS side information sent to the decoder is very small and does not exceed 31 bits. The idea is to start with one bit that specifies whether the size-N frame has been partitioned into two blocks of size N/2 each. If this bit is 1, it is followed by two bits that indicate whether any of these blocks has been partitioned into two blocks of size N/4 each. If either of the two bits is 1, they are followed by four bits that indicate which of the four size-N/4 blocks, if any, has been partitioned into two blocks of size N/8. The four bits may be followed by eight bits (for any blocks of size N/16) and then by another 16 bits (for any blocks of size N/32), for a total of at most 31 bits. As an example, the block structure of Figure 10.42c is described by the 31-bit BS 1|10|0100|00000000|0000000000000000. The simplest representation of BS is to always send the 31 bits to the decoder, but a small gain may be achieved by preceding BS with two “count” bits that specify its length. Here is how this can be done. Our BS consists of five fields of which the first is a single bit and must always exist. The remaining four fields are needed only if the the first field is 1. Thus, two count bits preceding the first bit of BS are sufficient to specify how many of the remaining four fields follow the first 10.13 MPEG-4 Audio Lossless Coding (ALS) 1025 field. For example, xx0 means any two count bits followed by a first field of 0, indicating no block switching, whereas 101|10|0100 indicates two fields (two bits and four bits) following the first bit of 1 (the count field is 102 = 2). Random Access. A practical, user-friendly audio compression algorithm should allow the user to skip forward and back to any desired point in the bitstream and decode the audio from that point. This feature, referred to as random access or seeking, is implemented in ALS by means of random-access frames. Normally, the first K audio samples of a frame are predicted by samples from the preceding frame, which prevents the decoder from jumping to the start of a frame and decoding it without first decoding the previous frame. A random-access frame is a special frame that can be decoded without decoding the preceding frame. Assuming that a frame corresponds to 50 ms of audio, if every 10th frame is generated by the encoder as a random-access frame, the decoder can skip to points in the audio with a precision of 500 ms, more than acceptable for normal audio listening, editing, and streaming applications. The size of a frame and the distance between random-access frames are user-controlled parameters. The latter can be from 1 to 255 frames. The only problem with random access in ALS is the prediction of the first K audio samples of a random-access frame. A Kth-order linear predictor normally predicts the first K samples of a frame using as many samples as needed from the previous frame, but in a random-access frame, the ALS predictor uses only samples from the current frame and has to predict, for example, the 4th sample from the first three samples of the frame. This is termed progressive prediction. The first sample cannot be predicted. The second sample is predicted from the first (first-order prediction), the third sample is predicted from the first two (second-order), and so on until sample K + 1 is predicted from its K predecessors in the same frame. Figure 10.43 summarizes this process. Part (a) shows a typical sequence of audio samples. Part (b) shows the result of continuous (non-progressive) prediction, where the first K samples of a random-access frame are not predicted (so the residues equal the original samples). Part (c) illustrates the results of progressive prediction. It shows how the first residue of the random-access frame is identical to the first sample, but successive residues get smaller and smaller. Long-Term Prediction. A pixel in a digital image is correlated with its near neighbors and an audio sample is similarly correlated with its neighbor samples, but there is a difference between correlation in images and in sound. Many musical instruments generate sound that consists of a fundamental frequency and higher harmonics that are multiples of that frequency. The concept of harmonics is easy to grasp when we consider a string instrument (but is also valid for wind and precussion instruments). A string is attached to the instrument at both ends. When plucked, the string tends to vibrate as shown in Figure 10.44a where it takes the form of half a wave. The fundamental frequency (or first harmonics) is therefore associated with a wavelength that is twice the length of the string. However, part of the time the string also vibrates as in part (b) of the figure, where its center is stationary. This creates the second harmonics frequency, whose associated wavelength equals the length of the string. If the string vibrates as in part (c) of the figure, it generates the third harmonics, with a wavelength that equals two-thirds of the string. In general, the wavelength . of the nth harmonics is related to the length L of the string by . = (2/n)L, which means that the frequency f of 1026 10. Audio Compression predicted fully predicted samples fully predicted samples samples fully predicted samples fully predicted samples non-predicted samples with reduced order random access point (a) (b) (c) Figure 10.43: Progressive Prediction in ALS. the nth harmonics is given by f = s/. = (s/2L)n, where s is the speed of the wave (notice that the speed depends on the medium of the string, but not on the wavelength or frequency). Thus, the frequencies of higher harmonics are integer multiples of the fundamental frequency. (a) (b) (c) Figure 10.44: Fundamental and Harmonic Frequencies. We normally don’t hear the harmonics as separate tones, mostly because they have increasingly lower amplitudes than the fundamental frequency, but also because our ear-brain system gets used to them from an early age and it combines them with the fundamental frequency. However, higher harmonics are normally present in sound generated by speech or by musical instruments, and they contribute to the richness of the audio we hear. A “pure” tone, with only a fundamental frequency and without any higher frequencies, can be generated by electronic devices, but it sounds artificial, thin, and uninteresting. Thus, harmonics exist in many audio streams and they contribute to the correlation between audio samples. However, it is easy to see that this correlation is long range. The frequency of the middle C musical note is 261 Hz. At a sampling rate of 48 kHz, a complete sine wave of middle C corresponds to 48,000/261 = 184 audio samples, and at a sampling rate of 192 kHz, a complete sine wave corresponds to 192,000/261 = 736 audio samples. Thus, there may be some correlation between two audio samples separated by hundreds of samples, and ALS tries to take advantage of this correlation. ALS employs a long-term predictor (LTP) that starts with a residue e(n) and computes a new residue ˜e(n) by subtracting from e(n) the contributions of five distant but related residues e(n - r + j). Parameter r is the lag. It has to be determined by the 10.13 MPEG-4 Audio Lossless Coding (ALS) 1027 encoder depending on the current frequency of the audio input and the sampling rate. Index j varies over five values. The rule of computation is ˜e(n) = e(n) -.. 2 j=-2 .r+j · e(n - r + j).. , where the five coefficients . are the quantized gain values. Both r and the five gain values have to be determined by the encoder (using an unspecified algorithm) and sent to the decoder as side information. The resulting long-term residue ˜e(n) is then encoded instead of the short-term residual e(n). [Notice that ˜e(n) is used instead of e(n) also for multi-channel prediction.] The decoder computes a short-term residue e(n) from a long-term residue ˜e(n) by means of e(n) = ˜e(n) +.. 2 j=-2 .r+j · e(n - r + j).. , and then employs e(n) and LPC to determine the next audio sample. Joint Channel Coding. ALS can support up to 216 audio channels, which is why it is important to exploit any possible correlations and dependencies between channels. Two schemes for multichannel coding are implemented as part of ALS and either can be used. 1. Difference Coding. Often, two audio channels x1(n) and x2(n) are known or suspected to be correlated. To take advantage of this, many compression methods compute the difference d(n) = x1(n) - x2(n) and encode either the pair [d(n), x1(n)] or the pair [d(n), x2(n)]. An even better approach is to use LPC to compute and encode residues for the three pairs 00 = [x1(n), x2(n)], 01 = [d(n), x1(n)], and 10 = [d(n), x2(n)] and select the smallest of the three residues. If this is done on a sample by sample basis, then the decoder has to be told which of the three pairs were selected for audio sample i. This information requires two bits, but only three of the four 2-bit combinations are used. Therefore, this side information should be compressed. We can consider it a sequence of trits (ternary digits), one trit for each audio sample. At the end of a block, the encoder tries to compress the sequence for the block with either run length or by counting the number of occurrences of 00, 01, and 10, and assigning them three appropriate Huffman codes. 2. Multichannel Coding. A volcano in a remote south sea island starts showing signs of an imminent eruption. Geologists rush to the site and place many seismometers on and around the mountain. The signals generated by the seismometers are correlated and serve as an example of many channels of correlated samples. Another example of many correlated channels is biomedical signals obtained from a patient by several measuring devices. The principle used by ALS to code many correlated channels is to select one channel as a reference r and encode it independently by predicting its samples, computing residues er(n), and encoding them. Any other channel c is coded by computing residues ec(n) and then making each residue smaller by subtracting from it a linear combination 1028 10. Audio Compression of three corresponding residues (three taps) er(n - 1), er(n), and er(n + 1) from the reference channel. The computation is ˆec(n) = ec(n) -.. 1 j=-1 .j · er(n + j).. , where the three gain coefficients .j are computed by solving the system of equations . = X-1 · y, (10.11) where . = (.-1, .0, .+1)T , X =.. er -1 T · er -1 er -1 T · er 0 er -1 T · er +1 er -1 T · er 0 er 0 T · er 0 er 0 T · er +1 er -1 T · er +1 er -1 T · er +1 er +1 T · er +1 .. , y = ecT · er -1, ecT · er 0, ecT · er +1, ec = ec(0), ec(1), ec(2), . . . , ec(N - 1)T , er -1 = er(0), er(1), er(2), . . . , er(N - 1)T , er +1 = er(1), er(2), er(3), . . . , er(N)T . (Where N is the frame size and T denotes a transpose.) The resulting residues ˆec(n) are then encoded and written on the bitstream. The decoder reconstructs the original residual signal by applying the reverse operation ec(n) = ˆec(n) +.. 1 j=-1 .j · er(n + j).. . There is also a mode of multichannel coding that employs six taps for long-range prediction. Residues ec(n) of a coding channel are computed and are then made smaller by subtracting a linear combination of six residues from the reference channel. These include the three residues er(n - 1), er(n), and er(n + 1) that correspond to ec(n) and another set of three consecutive residues at a distance of r from n, where the lag parameter r is to be estimated by the encoder by cross correlation between the coding channel and the reference channel. The computation is ˆec(n) = ec(n) -.. 1 j=-1 .j · er(n + j) + 1 j=-1 .r+j · er(n + r + j).. . The six gain parameters . can be obtained by minimizing the energy of the six subtracted residues, similar to Equation (10.11). The decoder, as usual, reconstructs the original 10.13 MPEG-4 Audio Lossless Coding (ALS) 1029 residues by the similar computation ec(n) = ˆec(n) +.. 1 j=-1 .j · er(n + j) + 1 j=-1 .r+j · er(n + r + j).. . Once the reconstructed residual signal ec(n) is obtained by the decoder, it is used by short-term and/or long-term linear prediction to obtain the reconstructed audio sample. Encoding the Residues. Two modes, simple and advanced, are available to the ALS encoder for encoding the residues e(n). In the simple mode, only Rice codes are used. The enocder has to determine the Rice parameter, and a sophisticated encoder selects the best parameter by computing the distribution of the residues. One option in the simple mode is to use the same Rice parameter for all the residues in a block. An alternative is to partition a block into four parts and use a different Rice parameter to encode the residues in each part. The Rice parameters have to be sent to the decoder as side information as shown in Figure 10.39. The advanced mode determines a value emax and encodes each residue e(n) according to its size relative to emax as shown in Figure 10.45. Residues that satisfy |e(n)| < emax are located in the central region of the residue distribution and are encoded by a special version of block Gilbert-Moore codes (BGMC, [Gilbert and Moore 59] and [Reznik 04]). Each residue is split into a most-significant part that is coded with BGMC and a least-significant part that is written in raw format (fixed-length) on the bitstream. Reference [Reznik 04] has more information on how the encoder can select reasonable values for emax and on the number of least-significant bits. Residues that satisfy |e(n)| . emax are located in the tails of the residue distribution. They are first recentered by et(n) = e(n)-emax and then encoded in Rice codes as in the simple mode. -4 -3 -2 -1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 1 Encoded with BGMC Encoded with Rice Encoded with Rice 2 3 4 Figure 10.45: Advanced Encoding of Residues. Floating-Point Audio Samples. It has been mentioned earlier that ALS supports audio signals in floating-point format, specifically, in the IEEE 32-bit format [IEEE754 85]. The encoder separates the sequence X of floating-point samples of a block into the sum of an integer sequence Y and a sequence Z of small floating-point 1030 10. Audio Compression numbers. The sequence Y of integer values is compressed in the normal manner, while the sequence Z is compressed by an algorithm termed masked Lempel-Ziv tool, a variant of LZW. The original sequence X of floating-point audio samples is separated in the form X = A.Y+Z, where X, Y, and Z are vectors, . denotes multiplication with rounding, and A is a common multiplier, a floating-point number in the interval [1, 2). The main problem is to select a common multiplier that will minimize the floating-point values in vector Z. The ALS standard proposes an approach to this problem, but does not guarantee that this approach is the best one. For a detailed discussion of this point as well as masked Lempel-Ziv tool, see [Liebchen et al. 05]. 10.14 MPEG-1/2 Audio Layers The video compression aspects of the MPEG-1 and MPEG-2 standards are described in Section 9.6. Its audio compression principles are discussed here. Readers are advised to read Sections 10.2 and 10.3 before trying to tackle this material. Some references for this section are [Brandenburg and Stoll 94], [Pan 95], [Rao and Hwang 96], and [Shlien 94]. The formal name of MPEG-1 is the international standard for moving picture video compression, IS 11172. It consists of five parts, of which part 3 [ISO/IEC 93] is the definition of the audio compression algorithm. Like other standards developed by the ITU and ISO, the document describing MPEG-1 has normative and informative sections. A normative section is part of the standard specification. It is intended for implementers, it is written in a precise language, and it should be strictly followed in implementing the standard on actual computer platforms. An informative section, on the other hand, illustrates concepts that are discussed elsewhere, explains the reasons that led to certain choices and decisions, and contains background material. An example of a normative section is the tables of various parameters and of the Huffman codes used in MPEG audio. An example of an informative section is the algorithm used by MPEG audio to implement a psychoacoustic model. MPEG does not require any particular algorithm, and an MPEG encoder can use any method to implement the model. This informative section simply describes various alternatives. The MPEG-1 and MPEG-2 (or in short, MPEG-1/2) audio standard specifies three compression methods called layers and designated I, II, and III. All three layers are part of the MPEG-1 standard. A movie compressed by MPEG-1 uses only one layer, and the layer number is specified in the compressed stream. Any of the layers can be used to compress an audio file without any video. The functional modules of the lower layers are also used by the higher layers, but the higher layers have additional features that result in better compression. An interesting aspect of the design of the standard is that the layers form a hierarchy in the sense that a layer-III decoder can also decode audio files compressed by layers I or II. The result of having three layers was an increasing popularity of layer III. The encoder is extremely complex, but it produces excellent compression, and this, combined with the fact that the decoder is much simpler, has produced in the late 1990s an explosion of what is popularly known as mp3 audio files. It is easy to legally and freely 10.14 MPEG-1/2 Audio Layers 1031 obtain a layer-III decoder and much music that is already encoded in layer III. So far, this has been a big success of the audio part of the MPEG project. The MPEG audio standard [ISO/IEC 93] starts with the normative description of the format of the compressed stream for each of the three layers. It follows with a normative description of the decoder. The description of the encoder (it is different for the three layers) and of two psychoacoustic models follows and is informative; any encoder that generates a correct compressed stream is a valid MPEG encoder. There are appendices (annexes) discussing related topics such as error protection. In contrast with MPEG video, where many information sources are available, there is relatively little in the technical literature about MPEG audio. In addition to the references elsewhere in this section, the reader is referred to the MPEG consortium [MPEG 00]. This site contains lists of other resources, and is updated from time to time. Another resource is the audio engineering society (AES). Most of the ideas and techniques used in the MPEG audio standard (and also in other audio compression methods) were originally published in the many conference proceedings of this organization. Unfortunately, these are not freely available and can be purchased from the AES. The history of MPEG-1 audio starts in December 1988, when a group of experts met for the first time in Hanover, Germany, to discuss the topic. During 1989, no fewer than 14 algorithms were proposed by these experts. They were eventually merged into four groups, known today as ASPEC, ATAC, MUSICAM, and SB-ADPCM. These were tested and compared in 1990, with results that led to the idea of adopting the three layers. Each of the three layers is a different compression method, and they increase in complexity (and in performance) from layer I to layer III. The first draft of the ISO MPEG-1 audio standard (standard 11172 part 3) was ready in December 1990. The three layers were implemented and tested during the first half of 1991, and the specifications of the first two layers were frozen in June 1991. Layer III went through further modifications and tests during 1991, and the complete MPEG audio standard was ready and was sent by the expert group to the ISO in December 1991, for its approval. It took the ISO/IEC almost a year to examine the standard, and it was finally approved in November 1992. When a movie is digitized, the audio part may often consist of two sound tracks (stereo sound), each sampled at 44.1 kHz with 16-bit samples. The audio data rate is therefore 2Χ44,100Χ16 = 1,411,200 bits/sec; close to 1.5 Mbits/sec. In addition to the 44.1 kHz sampling rate, the MPEG standard allows sampling rates of 32 kHz and 48 kHz. An important feature of MPEG audio is its compression ratio, which the standard specifies in advance! The standard calls for a compressed stream with one of several bitrates (Table 10.46) ranging from 32 to 224 kbps per audio channel (there normally are two channels, for stereo sound). Depending on the original sampling rate, these bitrates translate to compression factors of from 2.7 (low) to 24 (impressive)! The reason for specifying the bitrates of the compressed stream is that the stream also includes compressed data for the video and system parts. (These parts are mentioned in Section 9.6 and are not the same as the three audio layers.) The principle of MPEG audio compression is quantization. The values being quantized, however, are not the audio samples but numbers (called signals) taken from the frequency domain of the sound (this is discussed in the next paragraph). The fact that the compression ratio (or equivalently, the bitrate) is known to the encoder means that 1032 10. Audio Compression Bitrate (Kbps) Index Layer I Layer II Layer III 0000 free format free format free format 0001 32 32 32 0010 64 48 40 0011 96 56 48 0100 128 64 56 0101 160 80 64 0110 192 96 80 0111 224 112 96 1000 256 128 112 1001 288 160 128 1010 320 192 160 1011 352 224 192 1100 384 256 224 1101 416 320 256 1110 448 384 320 1111 forbidden forbidden forbidden Table 10.46: Bitrates in the Three Layers. the encoder knows at any time how many bits it can allocate to the quantized signals. Thus, the (adaptive) bit allocation algorithm is an important part of the encoder. This algorithm uses the known bitrate and the frequency spectrum of the most recent audio samples to determine the size of the quantized signals such that the quantization noise (the difference between an original signal and a quantized one) will be inaudible (i.e., will be below the masked threshold, a concept discussed in Section 10.3). The psychoacoustic models use the frequency of the sound that is being compressed, but the input stream consists of audio samples, not sound frequencies. The frequencies have to be computed from the samples. This is why the first step in MPEG audio encoding is a discrete Fourier transform, where a set of 512 consecutive audio samples is transformed to the frequency domain. Since the number of frequencies can be huge, they are grouped into 32 equal-width frequency subbands (layer III uses different numbers but the same principle). For each subband, a number is obtained that indicates the intensity of the sound at that subband’s frequency range. These numbers (called signals) are then quantized. The coarseness of the quantization in each subband is determined by the masking threshold in the subband and by the number of bits still available to the encoder. The masking threshold is computed for each subband using a psychoacoustic model. MPEG uses two psychoacoustic models to implement frequency masking and temporal masking. Each model describes how loud sound masks other sounds that happen to be close to it in frequency or in time. The model partitions the frequency range into 24 critical bands and specifies how masking effects apply within each band. The masking effects depend, of course, on the frequencies and amplitudes of the tones. When the sound is decompressed and played, the user (listener) may select any playback ampli- 10.14 MPEG-1/2 Audio Layers 1033 tude, which is why the psychoacoustic model has to be designed for the worst case. The masking effects also depend on the nature of the source of the sound being compressed. The source may be tone-like or noise-like. The two psychoacoustic models employed by MPEG are based on experimental work done by researchers over many years. The decoder must be fast, since it may have to decode the entire movie (video and audio) at real time, so it must be simple. As a result it does not use any psychoacoustic model or bit allocation algorithm. The compressed stream must therefore contain all the information that the decoder needs for dequantizing the signals. This information (the size of the quantized signals) must be written by the encoder on the compressed stream, and it constitutes overhead that should be subtracted from the number of remaining available bits. Figure 10.47 is a block diagram of the main components of the MPEG audio encoder and decoder. The ancillary data is user-definable and would normally consist of information related to specific applications. This data is optional. (a) (b) PCM audio input compressed output signal to mask ratio compressed output PCM audio decoded output stream formatting frequency domain mapping psychoacoustic model bit/noise allocation, quantizer, and coding ancillary data (optional) ancillary data (if included) frequency mapping unpacking to time frequency sample reconstruction Figure 10.47: MPEG Audio: (a) Encoder and (b) Decoder. 10.14.1 Frequency Domain Coding The first step in encoding the audio samples is to transform them from the time domain to the frequency domain. This is done by a bank of polyphase filters that transform the samples into 32 equal-width frequency subbands. The filters were designed to provide fast operation combined with good time and frequency resolutions. As a result, their design involved three compromises. The first compromise is the equal widths of the 32 frequency bands. This simplifies the filters but is in contrast to the behavior of the human auditory system, whose sensitivity is frequency-dependent. Ideally, the filters should divide the input into the critical bands discussed in Section 10.3. These bands are formed such that the perceived loudness of a given sound and its audibility in the presence of another, masking, sound are consistent within a critical band, but different across these bands. Unfortunately, each of the low-frequency subbands overlaps several critical bands, with the result that the bit allocation algorithm cannot optimize the number of bits allocated to the quantized 1034 10. Audio Compression signal in those subbands. When several critical bands are covered by a subband X, the bit allocation algorithm selects the critical band with the least noise masking and uses that critical band to compute the number of bits allocated to the quantized signals in subband X. The second compromise involves the inverse filter bank, the one used by the decoder. The original time-to-frequency transformation involves loss of information (even before any quantization). The inverse filter bank therefore receives data that is slightly bad, and uses it to perform the inverse frequency-to-time transformation, resulting in more distortions. Therefore, the design of the two filter banks (for direct and inverse transformations) had to use compromises to minimize this loss of information. The third compromise has to do with the individual filters. Adjacent filters should ideally pass different frequency ranges. In practice, they have considerable frequency overlap. Sound of a single, pure, frequency can therefore penetrate through two filters and produce signals (that are later quantized) in two of the 32 subbands instead of in just one subband. The polyphase filter bank uses (in addition to other intermediate data structures) a buffer X with room for 512 input samples. The buffer is a FIFO queue and always contains the most-recent 512 samples input. Figure 10.48 shows the five main steps of the polyphase filtering algorithm. 1. Shift in 32 new input samples into FIFO buffer X 2. Window samples: Zi = CiΧXi, for i = 0, 1, . . . , 511 3. Partial computation: Yi = 7 j=0 Zi+64j , for i = 0, 1, . . . , 63 4. Compute 32 signals: Si = 63 k=0Mi,kΧYk, for i = 0, 1, . . . , 31 5. Output the 32 subband signals Si Figure 10.48: Polyphase Filter Bank. The next 32 audio samples read from the input are shifted into the buffer. Thus, the buffer always holds the 512 most recent audio samples. The signals St[i] for the 32 subbands are computed by St[i] = 63 k=0 7 j=0 Mi,kC[k + 64j]ΧX[k + 64j], i = 0, . . . , 31. (10.12) The notation St[i] stands for the signal of subband i at time t. Vector C contains 512 coefficients of the analysis window and is fully specified by the standard. Some of the subband filter coefficients Ci are listed in Table 10.49a. M is the analysis matrix defined 10.14 MPEG-1/2 Audio Layers 1035 by Mi,k = cos (2i + 1)(k - 16). 64 , i = 0, . . . , 31, k = 0, . . . , 63. (10.13) Notice that the expression in parentheses in Equation (10.12) does not depend on i, while Mi,k in Equation (10.13) does not depend on j. (This matrix is a variant of MDCT, which is a modified version of the well-known DCT matrix.) This feature is a compromise that results in fewer arithmetic operations. In fact, the 32 signals St[i] are computed by only 512 + 32 Χ 64 = 2560 multiplications and 64 Χ 7 + 32 Χ 63 = 2464 additions, which come to about 80 multiplications and 80 additions per signal. Another point worth mentioning is the decimation of samples (Section 8.7). The entire filter bank produces 32 output signals for 32 input samples. Since each of the 32 filters produces 32 signals, its output has to be decimated, retaining only one signal per filter. Figure 10.51a,b illustrates graphically the operations performed by the encoder and decoder during the polyphase filtering step. Part (a) of the figure shows how the X buffer holds 64 segments of 32 audio samples each. The buffer is shifted one segment to the right before the next segment of 32 new samples is read from the input stream and is entered on the left. After multiplying the X buffer by the coefficients of the C window, the products are moved to the Z vector. The contents of this vector are partitioned into segments of 64 numbers each, and the segments are added. The result is vector Y, which is multiplied by the MDCT matrix, to produce the final vector of 32 subband signals. Part (b) of the figure illustrates the operations performed by the decoder. A group of 32 subband signals is multiplied by the IMDCT matrix Ni,k to produce the V vector, consisting of two segments of 32 values each. The two segments are shifted into the V FIFO buffer from the left. The V buffer has room for the last 16 V vectors (i.e., 16Χ64, or 1024, values). A new 512-entry U vector is created from 32 alternate segments in the V buffer, as shown. The U vector is multiplied by the 512 coefficients Di of the synthesis window (similar to the Ci coefficients of the analysis window used by the encoder), to create the W vector. Some of the Di coefficients are listed in Table 10.49b. This vector is divided into 16 segments of 32 values each and the segments added. The result is 32 reconstructed audio samples. Figure 10.50 is a flowchart illustrating this process. The IMDCT synthesis matrix Ni,k is given by Ni,k = cos (2k + 1)(i + 16). 64 , i = 0, . . . , 63, k = 0, . . . , 31. The subband signals computed by the filtering stage of the encoder are collected and packaged into frames containing 384 signals (in layer I) or 1152 signals (in layers II and III) each. The signals in a frame are then scaled and quantized according to the psychoacoustic model used by the encoder and the bit allocation algorithm. The quantized values, together with the scale factors and quantization information (the number of quantization levels in each subband) are written on the compressed stream (layer III also uses Huffman codes to encode the quantized values even further). 10.14.2 Format of Compressed Data In layer I, each frame consists of 12 signals per subband, for a total of 12Χ32 = 384 signals. In layers II and III, a frame contains 36 signals per subband, for a total of 1036 10. Audio Compression i Ci i Ci i Ci 0 0.000000000 252 -0.035435200 506 -0.000000477 1 -0.000000477 253 -0.035586357 507 -0.000000477 2 -0.000000477 254 -0.035694122 508 -0.000000477 3 -0.000000477 255 -0.035758972 509 -0.000000477 4 -0.000000477 256 -0.035780907 510 -0.000000477 5 -0.000000477 257 -0.035758972 511 -0.000000477 ... ... (a) i Di i Di i Di 0 0.000000000 252 -1.133926392 506 0.000015259 1 -0.000015259 253 -1.138763428 507 0.000015259 2 -0.000015259 254 -1.142211914 508 0.000015259 3 -0.000015259 255 -1.144287109 509 0.000015259 4 -0.000015259 256 -1.144989014 510 0.000015259 5 -0.000015259 257 -1.144287109 511 0.000015259 ... ... (b) Table 10.49: Coefficients of (a) the Analysis and (b) the Synthesis Window. 1. Input 32 new signals Si, for i = 0, . . . , 31 2. Shift* the FIFO buffer Vi = Vi-64, i = 1023, . . . , 64 3. Multiply Vi = 31 k=0 NikSk, i = 0, . . . , 63 4. Construct 512 values in vector U for i = 0 to 7 do for j = 0 to 31 do Uj+64i = Vj+128i U32+j+64i = V96+j+128i 5. Construct W. For i = 0 to 511 do, Wi = UiΧDi 6. Compute 32 samples. For j = 0 to 31 do, Sj = 15 i=0Wj+32i 7. Output the 32 audio samples Sj * V is initialized to all zeros at startup. Figure 10.50: Reconstructing Audio Samples. 10.14 MPEG-1/2 Audio Layers 1037 (a) (b) + signals MDCT subband 64 64 vector Y temporary vector Z coefficient vector C fifo buffer X 32 samples read from input stream + + = vector V reconstructed + + + + + = samples + 0 0 0 63 31 0 31 0 63 0 31 0 31 511 511 vector U 0 511 window D 0 511 vector W 0 511 V fifo buffer 16 V vectors =1024 samples Χ Χ 64Χ32 IMDCT 64Χ32 Figure 10.51: MPEG Audio: (a) Encoder and (b) Decoder. 1038 10. Audio Compression 1152 signals. The signals in the frame are quantized (this is the important step where compression is achieved) and written on the compressed stream together with other information. Each frame written on the output starts with a 32-bit header whose format is identical for the three layers. The header contains a synchronization code (12 1-bits) and 20 bits of coding parameters listed below. If error protection is used, the header is immediately followed by a 16-bit CRC check word (Section 6.32) that uses the generating polynomial CRC16(x) = x16+x15+x2+1. Next comes a frame of the quantized signals, followed by an (optional) ancillary data block. The formats of the last two items depend on the layer. In layer I, the CRC 16-bit code is computed from the last 16 bits (bits 16–31) of the header and from the bit allocation information. In layer II, the CRC code is computed from these bits plus the scalefactor selection information. In layer III, the CRC code is computed from the last 16 bits of the header and also from either bits 0–135 of the audio data (in single-channel mode) or bits 0–255 of the audio data (in the other three modes, see field 8 below). The synchronization code is used by the decoder to verify that what it is reading is, in fact, the header. This code is a string of 12 bits of 1, so the entire format of the compressed stream had to be designed to avoid an accidental occurrence of a string of twelve 1’s elsewhere. The remaining 20 bits of the header are divided into 12 fields as follows: Field 1. An ID bit whose value is 1 (this indicates that MPEG is used). A value of 0 is reserved and is currently unused. Field 2. Two bits to indicate the layer number. Valid values are 11—layer I, 10— layer II, and 01—layer III. The value 00 is reserved. Field 3. An error protection indicator bit. A value of 0 indicates that redundancy has been added to the compressed stream to help in error detection. Field 4. Four bits to indicate the bitrate (Table 10.46). A zero index indicates a “fixed” bitrate, where a frame may contain an extra slot, depending on the padding bit (field 6). Field 5. Two bits to indicate one of three sampling frequencies. The three values are 00—44.1 kHz, 01—48 kHz, and 10—32 kHz. The value 11 is reserved. Field 6. One bit to indicate whether padding is used. Padding may add a slot (slots are discussed in Section 10.14.3) to the compressed stream after a certain number of frames, to make sure that the total size of the frames either equals or is slightly less than the sum current frame first frame frame-sizeΧbitrate sampling frequency, where frame-size is 384 signals for layer I and 1152 signals for layers II and III. The following algorithm may be used by the encoder to determine whether or not padding is necessary. for first audio frame: rest:=0; padding:=no; 10.14 MPEG-1/2 Audio Layers 1039 for each subsequent audio frame: if layer=I then dif:=(12 Χ bitrate) modulo (sampling-frequency) else dif:=(144 Χ bitrate) modulo (sampling-frequency); rest:=rest-dif; if rest<0 then padding:=yes; rest:=rest+(sampling-frequency) else padding:=no; This algorithm has a simple interpretation. A frame is divided into N or N + 1 slots, where N depends on the layer. For layer I, N is given by N = 12Χ bitrate sampling frequency. For layers II and III, it is given by N = 144Χ bitrate sampling frequency. If this does not produce an integer, the result is truncated and padding is used. Padding is also mentioned in Section 10.14.3 Field 7. One bit for private use of the encoder. This bit will not be used by ISO/IEC in the future. Field 8. A two-bit stereo mode field. Values are 00—stereo, 01—joint-stereo (intensity-stereo and/or ms-stereo), 10—dual-channel, and 11—single-channel. Stereo information is encoded in one of four modes: stereo, dual channel, joint stereo, and ms-stereo. In the first two modes, samples from the two stereo channels are compressed and written on the output. The encoder does not check for any correlations between the two. The stereo mode is used to compress the left and right stereo channels, while the dual channel mode is used to compress different sets of audio samples, such as a bilingual broadcast. The joint stereo mode exploits redundancies between the left and right channels, since many times they are identical, similar, or differ by a small time lag. The ms-stereo mode (“ms” stands for “middle-side”) is a special case of joint stereo, where two signals, a middle value Mi and a side value Si, are encoded instead of the left and right audio channels Li and Ri. The middle-side values are computed by the following sum and difference Li = Mi + Si .2 , and Ri = Mi - Si .2 . Field 9. A two-bit mode extension field. This is used in the joint-stereo mode. In layers I and II the bits indicate which subbands are in intensity-stereo. All other subbands are coded in “stereo” mode. The four values are: 00—subbands 4–31 in intensity-stereo, bound=4. 01—subbands 8–31 in intensity-stereo, bound=8. 1040 10. Audio Compression 10—subbands 12–31 in intensity-stereo, bound=12. 11—subbands 16–31 in intensity-stereo, bound=16. In layer III these bits indicate which type of joint stereo coding method is applied. The values are: 00—intensity-stereo off, ms-stereo off. 01—intensity-stereo on, ms-stereo off. 10—intensity-stereo off, ms-stereo on. 11—intensity-stereo on, ms-stereo on. Field 10. Copyright bit. If the compressed stream is copyright protected, this bit should be 1. Field 11. One bit indicating original/copy. A value of 1 indicates an original compressed stream. Field 12. A 2-bit emphasis field. This indicates the type of de-emphasis that is used. The values are 00—none, 01—50/15 microseconds, 10—reserved, and 11 indicates CCITT J.17 de-emphasis. Layer I: The 12 signals of each subband are scaled such that the largest one becomes 1 (or very close to 1, but not greater than 1). The psychoacoustic model and the bit allocation algorithm are invoked to determine the number of bits allocated to the quantization of each scaled signal in each subband (or, equivalently, the number of quantization levels). The scaled signals are then quantized. The number of quantization levels, the scale factors, and the 384 quantized signals are then placed in their areas in the frame, which is written on the output. Each bit allocation item is four bits, each scale factor is six bits, and each quantized sample occupies between two and 15 bits in the frame. The number l of quantization levels and the number q of bits per quantized value are related by 2q = l. The bit allocation algorithm uses tables to determine q for each of the 32 subbands. The 32 (q - 1) values are then written, as 4-bit numbers, on the frame, for the use of the decoder. Thus, the 4-bit value 0000 read by the decoder from the frame for subband s indicates to the decoder that the 12 signals of s have been quantized coarsely to one bit each, while the value 1110 implies that the 16-bit signals have been finely quantized to 15 bits each. The value 1111 is not used, to avoid conflict with the synchronization code. The encoder decides many times to quantize all the 12 signals of a subband s to zero, and this is indicated in the frame by a different code. In such a case, the 4-bit value input from the frame for s is ignored by the decoder.  Exercise 10.8: Explain why the encoder may decide to quantize 12 consecutive signals of subband s to zero. The decoder multiplies the dequantized signal values by the scale factors found in the frame. There is one scale factor for the 12 signals of a subband. This scale factor is selected by the encoder from a table of 63 scale factors specified by the standard (Table 10.52). The scale factors in the table increase by a factor of .3 2.  Exercise 10.9: How does this increase translate to an increase in the decibel level? Quantization is performed by the following simple rule: If the bit allocation algorithm allocates b bits to each quantized value, then the number n of quantization levels is determined by b = log2(n + 1), or 2b = n + 1. The value b = 3, for example, results 10.14 MPEG-1/2 Audio Layers 1041 Index scalefactor 0 2.00000000000000 1 1.58740105196820 2 1.25992104989487 3 1.00000000000000 4 0.79370052598410 5 0.62996052494744 6 0.50000000000000 7 0.39685026299205 8 0.31498026247372 9 0.25000000000000 10 0.19842513149602 11 0.15749013123686 12 0.12500000000000 13 0.09921256574801 14 0.07874506561843 15 0.06250000000000 16 0.04960628287401 17 0.03937253280921 18 0.03125000000000 19 0.02480314143700 20 0.01968626640461 Index scalefactor 21 0.01562500000000 22 0.01240157071850 23 0.00984313320230 24 0.00781250000000 25 0.00620078535925 26 0.00492156660115 27 0.00390625000000 28 0.00310039267963 29 0.00246078330058 30 0.00195312500000 31 0.00155019633981 32 0.00123039165029 33 0.00097656250000 34 0.00077509816991 35 0.00061519582514 36 0.00048828125000 37 0.00038754908495 38 0.00030759791257 39 0.00024414062500 40 0.00019377454248 41 0.00015379895629 Index scalefactor 42 0.00012207031250 43 0.00009688727124 44 0.00007689947814 45 0.00006103515625 46 0.00004844363562 47 0.00003844973907 48 0.00003051757813 49 0.00002422181781 50 0.00001922486954 51 0.00001525878906 52 0.00001211090890 53 0.00000961243477 54 0.00000762939453 55 0.00000605545445 56 0.00000480621738 57 0.00000381469727 58 0.00000302772723 59 0.00000240310869 60 0.00000190734863 61 0.00000151386361 62 0.00000120155435 Table 10.52: Layers I and II Scale Factors. in n = 7 quantization values. Figure 10.53a,b shows examples of such quantization. The input signals being quantized have already been scaled to the interval [-1, +1], and the quantization is midtread. For example, all input values in the range [-1/7,-3/7] are quantized to 010 (dashed lines in the figure). Dequantization is also simple. The quantized value 010 is always dequantized to -2/7. Notice that quantized values range from 0 to n - 1. The value n = 11. . . 1 is never used, in order to avoid a conflict with the synchronization code. Table 10.54 shows that the bit allocation algorithm for layer I can select the size of the quantized signals to be between 0 and 15 bits. For each of these sizes, the table lists the 4-bit allocation code that the encoder writes on the frame for the decoder’s use, the number q of quantization levels (between 0 and 215-1 = 32,767, and the signal-to-noise ratio in decibels. In practice, the encoder scales a signal Si by a scale factor scf that is determined from Table 10.52, and quantizes the result by computing Sqi = A Si scf + B ))))N , where A and B are the constants listed in Table 10.55 (the three entries flagged by asterisks are used by layer II but not by layer I) and N is the number of bits needed to encode the number of quantization levels. (The vertical bar with subscript N means: Take the N most-significant bits.) In order to avoid conflicts with the synchronization 1042 10. Audio Compression (a) (b) input signal (scaled) quantized value quantized input dequantized value 000 001 010 011 100 101 110 7 -7 7 -6 7 -6 7 -5 7 -4 7 -4 7 -3 7 -2 7 -2 7 -1 7 0 1 000 001 010 011 100 101 110 0 7 2 7 2 7 3 7 4 7 4 7 5 7 6 7 6 7 7 Figure 10.53: Examples of (a) Quantizer and (b) Dequantizer. code, the most-significant bit of the quantized value Sqi is inverted before that quantity is written on the output. Layer II Format: This is an extension of the basic method outlined for layer I. Each frame now consists of 36 signals per subband, and both the bit allocation data and the scale factor information are coded more efficiently. Also, quantization can be much finer and can have up to 216 - 1 = 65,535 levels. A frame is divided into three parts, numbered 0, 1, and 2. Each part resembles a layer-I frame and contains 12 signals per subband. The bit allocation data is common to the three parts, but the information for the scale factors is organized differently. It can be common to all three parts, it can apply to just two of the three parts, or it can be specified for each part separately. This information consists of a 2-bit scale factor selection information (sfsi). This number indicates whether one, two, or three scale factors per subband are written in the frame, and how they are applied. The bit allocation section of the frame is encoded more efficiently by limiting the choice of quantization levels for the higher subbands and the lower bitrates. Instead of the four bits per subband used by layer I to specify the bit allocation choice, the number of bits used by layer II varies from 0 to 4 depending on the subband number. The MPEG 10.14 MPEG-1/2 Audio Layers 1043 bit 4-bit number alloc code of levels SNR (dB) 0 0000 0 0.00 2 0001 3 7.00 3 0010 7 16.00 4 0011 15 25.28 5 0100 31 31.59 6 0101 63 37.75 7 0110 127 43.84 8 0111 255 49.89 9 1000 511 55.93 10 1001 1023 61.96 11 1010 2047 67.98 12 1011 4095 74.01 13 1100 8191 80.03 14 1101 16383 86.05 15 1110 32767 92.01 invalid 1111 Table 10.54: Bit Allocation and Quantization in Layer I. number of levels A B 3 0.750000000 -0.250000000 5* 0.625000000 -0.375000000 7 0.875000000 -0.125000000 9* 0.562500000 -0.437500000 15 0.937500000 -0.062500000 31 0.968750000 -0.031250000 63 0.984375000 -0.015625000 127 0.992187500 -0.007812500 255 0.996093750 -0.003906250 511 0.998046875 -0.001953125 1023 0.999023438 -0.000976563 2047 0.999511719 -0.000488281 4095 0.999755859 -0.000244141 8191 0.999877930 -0.000122070 16383 0.999938965 -0.000061035 32767 0.999969482 -0.000030518 65535* 0.999984741 -0.000015259 Table 10.55: Quantization Coefficients in Layers I and II. standard contains a table that the encoder searches by subband number and bitrate to find the number of bits. 1044 10. Audio Compression Quantization is similar to layer I, except that layer II can sometimes pack three consecutive quantized values in a single codeword. This can be done when the number of quantization levels is a power of 2, and it reduces the number of wasted bits. 10.14.3 Encoding: Layers I and II The MPEG standard specifies a table of scale factors. For each subband, the encoder compares the largest of the 12 signals to the values in the table, finds the next largest value, and uses the table index of that value to determine the scale factor for the 12 signals. In layer II, the encoder determines three scale factors for each subband, one for each of the three parts. It calculates the difference between the first two and the last two and uses the two differences to decide whether to encode one, two, or all three scale factors. This process is described in Section 10.14.4. The bit allocation information in layer II uses 2–4 bits. The scale factor select information (sfsi) is two bits. The scale factor itself uses six bits. Each quantized signal uses 2–16 bits, and there may be ancillary data. The standard describes two psychoacoustic models. Each produces a quantity called the signal to mask ratio (SMR) for each subband. The bit allocation algorithm uses this SMR and the SNR from Table 10.54 to compute the mask to noise ratio (MNR) as the difference MNR = SMR - SNR dB. The MNR indicates the discrepancy between waveform error and perceptual measurement, and the idea is that the subband signals can be compressed as much as the MNR. As a result, each iteration of the bit allocation loop determines the minimum MNR of all the subbands. The basic principle of bit allocation is to minimize the MNR over a frame while using not more than the number of bits Bf available for the frame. In each iteration, the algorithm computes the number of bits Bf available to encode a frame. This is computed from the sample rate (number of samples input per second, normally 2Χ44,100) and the bitrate (number of bits written on the compressed output per second). The calculation is frames/sec = samples/second samples/frame, Bf = bits/second frames/second = bitrate Χ samples per frame sampling rate . Thus, Bf is measured in bits/frame. The frame header occupies 32 bits, and the CRC, if used, requires 16 bits. The bit allocation data is four bits per subband. If ancillary data are used, its size is also determined. These amounts are subtracted from the value of Bf computed earlier. The main step of the bit allocation algorithm is to maximize the minimum MNR for all subbands by assigning the remaining bits to the scale factors and the quantized signals. In layer I, the scale factors take six bits per subband, but in layer II there are several ways to encode them (Section 10.14.4). The main bit allocation step is a loop. It starts with allocating zero bits to each of the subbands. If the algorithm assigns zero bits to a particular subband, then no bits will be needed for the scale factors and the quantized signals. Otherwise, the number of bits assigned depends on the layer number and on the number of scale factors (1, 2, or 3) encoded for the subband. The algorithm computes the SNR and MNR for each 10.14 MPEG-1/2 Audio Layers 1045 subband and searches for the subband with the lowest MNR whose bit allocation has not yet reached its maximum limit. The bit allocation for that subband is incremented by one level, and the number of extra bits needed is subtracted from the balance of available bits. This is repeated until the balance of available bits reaches zero or until all the subbands have reached their maximum limit. The format of the compressed output is shown in Figure 10.56. Each frame consists of between two and four parts. The frame is organized in slots, where a slot size is 32 bits in layer I, and 8 bits in layers II and III. Thus, the number of slots is Bf /32 or Bf /8. If the last slot is not full, it is padded with zeros. sequence Frame Frame Frame 1 Frame 2 Frame n header CRC data ancillary Figure 10.56: Format of Compressed Output. The relation between sampling rate, frame size, and number of slots is easy to visualize. Typical layer I parameters are (1) a sampling rate of 48,000 samples/sec, (2) a bitrate of 64,000 bits/sec, and (3) 384 quantized signals per frame. The decoder has to decode 48,000/384 = 125 frames per second. Thus, each frame has to be decoded in 8 ms. In order to output 125 frames in 64,000 bits, each frame must have Bf = 512 bits available to encode it. The number of slots per frame is thus 512/32 = 16. A similar example assumes (1) a sampling rate of 32,000 samples/sec, (2) a bitrate of 64,000 bits/sec, and (3) 384 quantized signals per frame. The decoder has to decode 32,000/384 = 83.33 frames per second. Thus, each frame has to be decoded in 12 ms. In order to output 83.33 frames in 64000 bits, each frame must have Bf = 768 bits available to encode it. Thus, the number of slots per frame is 768/32 = 24.  Exercise 10.10: Do the same calculation for a sampling rate of 44,100 samples/s. 10.14.4 Encoding: Layer II A frame in layer II consists of 36 subband signals, organized in 12 granules as shown in Figure 10.57. Three scale factors scf1, scf2, and scf3 are computed for each subband, one scale factor for each group of 12 signals. This is done in six steps as follows: Step 1: The maximum of the absolute values of these 12 signals is determined. Step 2: This maximum is compared with the values in column “scalefactors” of Table 10.52, and the smallest entry that is greater than the maximum is noted. Step 3: The value in column “Index” of that entry is denoted by scfi. Step 4: After repeating steps 1–3 for i = 1, 2, 3, two differences, D1 = scf1 - scf2 and D2 = scf2 - scf3, are computed. 1046 10. Audio Compression Step 5: Two “class” values, for D1 and D2, are determined by Classi =.. .... ... 1, if Di . -3, 2, if -3 < Di < 0, 3, if Di = 0, 4, if 0 < Di < 3, 5, if Di . 3. Step 6: Depending on the two classes, three scale factors are determined from Table 10.58, column “s. fact. used.” The values 1, 2, and 3 in this column stand for the first, second, and third scale factors, respectively, within a frame. The value 4 means the maximum of the three scale factors. The column labeled “trans. patt.” indicates those scale factors that are actually written on the compressed stream. Table 10.59 lists the four values of the 2-bit scale factor select information (scfsi). As an example, suppose that the two differences D1 and D2 between the three scale factors A, B, and C of the three parts of a certain subband are classified as (1, 1). Table 10.58 shows that the transmission factor is 123 and the select information is 0. The top rule in Table 10.59 shows that scfsi of 0 means that each of the three scale factors is encoded separately in the output stream as a 6-bit number. (This rule is used by both encoder and decoder.) The three scale factors occupy, in this case, 18 bits, so there is no savings compared to layer I encoding. We next assume that the two differences are classified as (1, 3). Table 10.58 shows that the scale factor is 122 and the select information is 3. The bottom rule in Table 10.59 shows that a scfsi of 3 means that only scale factors A and B need be encoded (since B = C), occupying just 12 bits. The decoder assigns the first six bits as the value of A, and the following six bits as the values of both B and C. Thus, the redundancy in the scale factors has been translated to a small savings.  Exercise 10.11: How are the three scale factors encoded when the two differences are classified as (3, 2)? Quantization in layer II is similar to that of layer I with the difference that the two constants C and D of Table 10.60 are used instead of the constants A and B. Another difference is the use of grouping. Table 10.60 shows that grouping is required when the number of quantization levels is 3, 5, or 9. In any of these cases, three signals are combined into one codeword, which is then quantized. The decoder reconstructs the three signals s(1), s(2), and s(3) from the codeword w by w = 0, for i = 0 to 2, s(i) = w mod nos, w = w χ nos, where “mod” is the modulo function, “χ” denotes integer division, and “nos” denotes the number of quantization steps. Figure 10.61a,b shows the format of a frame in both layers I and II. In layer I (part (a) of the figure), there are 384 signals per frame. Assuming a sampling rate of 48,000 10.14 MPEG-1/2 Audio Layers 1047 32 subbands granules 1 granule 12 3Χ12Χ32=1,152 samples Figure 10.57: Organization of Layer II Subband Signals. class class s. fact trans. s. fact. class class s. fact trans. s. fact. 1 2 used patt. select 1 2 used patt. select 1 1 123 123 0 3 4 333 3 2 1 2 122 12 3 3 5 113 13 1 1 3 122 12 3 4 1 222 2 2 1 4 133 13 3 4 2 222 2 2 1 5 123 123 0 4 3 222 2 2 2 1 113 13 1 4 4 333 3 2 2 2 111 1 2 4 5 123 123 0 2 3 111 1 2 5 1 123 123 0 2 4 444 4 2 5 2 122 12 3 2 5 113 13 1 5 3 122 12 3 3 1 111 1 2 5 4 133 13 3 3 2 111 1 2 5 5 123 123 0 3 3 111 1 2 Table 10.58: Layer II Scale Factor Transmission Patterns. # of coded decoding scsfi scale factors scale factor 0 (00) 3 scf1, scf2, scf3 1 (01) 2 1st is scf1 and scf2 2nd is scf3 2 (10) 1 one scale factor 3 (11) 2 1st is scf1 2nd is scf2 and scf3 Table 10.59: Layer II Scale Factor Select Information. 1048 10. Audio Compression number of steps C D grouping samples/code bits/code 3 1.3333333333 0.5000000000 yes 3 5 5 1.6000000000 0.5000000000 yes 3 7 7 1.1428571428 0.2500000000 no 1 3 9 1.7777777777 0.5000000000 yes 3 10 15 1.0666666666 0.1250000000 no 1 4 31 1.0322580645 0.0625000000 no 1 5 63 1.0158730158 0.0312500000 no 1 6 127 1.0078740157 0.0156250000 no 1 7 255 1.0039215686 0.0078125000 no 1 8 511 1.0019569471 0.0039062500 no 1 9 1023 1.0009775171 0.0019531250 no 1 10 2047 1.0004885197 0.0009765625 no 1 11 4095 1.0002442002 0.0004882812 no 1 12 8191 1.0001220852 0.0002441406 no 1 13 16383 1.0000610388 0.0001220703 no 1 14 32767 1.0000305185 0.0000610351 no 1 15 65535 1.0000152590 0.0000305175 no 1 16 Table 10.60: Quantization Classes for Layer II. (a) (b) subband signals scale factors AD 6 bit ancillary data ancillary data unspecified length unspecified length 12 bit header 16 bit CRC synch 20 bit system allocation bit 4 bit scfsi subband signals scale factors AD 6 bit 12 granules of 3 signals each 12 bit header 16 bit CRC synch 20 bit system allocation bit low subbands 2 bit 4 bit med subbands 3 bit high subbands 2 bit ABC A C AB A Figure 10.61: Organization of Layers I and II Output Frames. 10.14 MPEG-1/2 Audio Layers 1049 samples/sec and a bitrate of 64,000 bits/sec, such a frame must be fully decoded in 80 ms, for a decoding rate of 125 frames/sec.  Exercise 10.12: What is a typical frame rate for layer II? 10.14.5 Psychoacoustic Models The task of a psychoacoustic model is to make it possible for the encoder to easily decide how much quantization noise to allow in each subband. This information is then used by the bit allocation algorithm, together with the number of available bits, to determine the number of quantization levels for each subband. The MPEG audio standard specifies two psychoacoustic models. Either model can be used with any layer, but only model II generates the specific information needed by layer III. In practice, model I is the only one used in layers I and II. Layer III can use either model, but it achieves better results when using model II. The MPEG standard allows considerable freedom in the way the models are implemented. The sophistication of the model that is actually implemented in a given MPEG audio encoder depends on the desired compression factor. For consumer applications, where large compression factors are not critical, the psychoacoustic model can be completely eliminated. In such a case, the bit allocation algorithm does not use an SMR (signal to mask ratio). It simply assigns bits to the subband with the minimum SNR (signal to noise ratio). A complete description of the models is outside the scope of this book and can be found in the text of the MPEG audio standard [ISO/IEC 93], pp. 109–139. The main steps of the two models are as follows: 1. A Fourier transform is used to convert the original audio samples to their frequency domain. This transform is separate and different from the polyphase filters because the models need finer frequency resolution in order to accurately determine the masking threshold. 2. The resulting frequencies are grouped into critical bands, not into the same 32 subbands used by the main part of the encoder. 3. The spectral values of the critical bands are separated into tonal (sinusoid-like) and nontonal (noise-like) components. 4. Before the noise masking thresholds for the different critical bands can be determined, the model applies a masking function to the signals in the different critical bands. This function has been determined empirically, by experimentation. 5. The model computes a masking threshold for each subband. 6. The SMR (signal to mask ratio) is calculated for each subband. It is the signal energy in the subband divided by the minimum masking threshold for the subband. The set of 32 SMRs, one per subband, constitutes the output of the model. 10.14.6 Encoding: Layer III Layer III employs a much more refined and complex algorithm than the first two layers. This is reflected in the compression factors, which are much higher. The difference between layer III and layers I and II starts at the very first step, filtering. The same polyphase filter bank (Table 10.49a) is used, but it is followed by a modified version of the discrete cosine transform. The MDCT corrects some of the errors introduced 1050 10. Audio Compression by the polyphase filters and also subdivides the subbands to bring them closer to the critical bands. The layer III decoder has to use the inverse MDCT, so it has to work harder. The MDCT can be performed on either a short block of 12 samples (resulting in six transform coefficients) or a long block of 36 samples (resulting in 18 transform coefficients). Regardless of the block size chosen, consecutive blocks transformed by the MDCT have considerable overlap, as shown in Figure 10.62. In this figure, the blocks are shown above the thick line, and the resulting groups of 18 or 6 coefficients are below the line. The long blocks produce better frequency spectrum for stationary sound (sound where adjacent samples don’t differ much), while the short blocks are preferable when the sound varies often. The MDCT uses n input samples xk (where n is either 36 or 12) to obtain n/2 (i.e., 18 or 6) transform coefficients Si. The transform and its inverse are given by Si = n-1 k=0 xk cos  . 2n 2k +1+ n 2 (2i + 1), i = 0, 1, . . . , n 2 - 1, (10.14) xk = n/2-1 i=0 Si cos  . 2n 2k +1+ n 2 (2i + 1), k = 0, 1, . . . , n - 1. (10.15) 1 3 5 7 9 11 13 15 17 1 3 5 7 9 11 13 15 17 1 3 5 7 9 11 13 15 17 19 21 23 1 3 5 1 3 5 1 3 5 1 3 5 1 3 5 1 3 5 1 3 5 7 9 11 13 15 17 1 3 5 7 9 11 13 15 17 36 samples 18 samples 6 samples 24 new samples 36 samples 36 samples 36 samples 36 samples 18 new samples 18 samples 18 samples 18 samples 18 new samples 18 new samples 18 new samples 18 samples 18 samples 18 samples 18 samples 12 samples 6 6 12 samples 6 6 12 samples 6 6 12 samples 6 6 12 samples 6 6 12 samples 6 6 6 new samples 6 new samples 6 new samples 6 new samples 6 new samples Figure 10.62: Overlapping MDCT Windows. The size of a short block is one-third that of a long block, so they can be mixed. When a frame is constructed, the MDCT can use all long blocks, all short blocks (three times as many), or long blocks for the two lowest-frequency subbands and short blocks for the remaining 30 subbands. This is a compromise where the long blocks provide finer frequency resolution for the lower frequencies, where it is most useful, and the short blocks maintain better time resolution for the high frequencies. Since the MDCT provides better frequency resolution, it has to result in poorer time resolution because of the uncertainty principle (Section 8.3). What happens in practice is that the quantization of the MDCT coefficients causes errors that are spread 10.14 MPEG-1/2 Audio Layers 1051 over time and cause audible distortions that manifest themselves as preechoes (read: “pre-echoes”). The psychoacoustic model used by layer III has extra features that detect conditions for preechoes. In such cases, layer III uses a complex bit allocation algorithm that borrows bits from the pool of available bits in order to temporarily increase the number of quantization levels and thus reduce preechoes. Layer III can also switch to short MDCT blocks, thereby reducing the size of the time window, if it “suspects” that conditions are favorable for preechoes. (The layer-III psychoacoustic model calculates a quantity called “psychoacoustic entropy” (PE), and the layer-III encoder “suspects” that conditions are favorable for preechoes if PE > 1800.) The MDCT coefficients go through some processing to remove artifacts caused by the frequency overlap of the 32 subbands. This is called aliasing reduction. Only the long blocks are sent to the aliasing reduction procedure. The MDCT uses 36 input samples to compute 18 coefficients, and the aliasing reduction procedure uses a butterfly operation between two sets of 18 coefficients. This operation is illustrated graphically in Figure 10.63a with a C-language code listed in Figure 10.63b. Index i in the figure is the distance from the last line of the previous block to the first line of the current block. Eight butterflies are computed, with different values of the weights csi and cai that are given by csi = 1 1 + c2i , cai = ci 1 + c2i , i= 0, 1, . . . , 7. The eight ci values specified by the standard are -0.6, -0.535, -0.33, -0.185, -0.095, -0.041, -0.0142, and -0.0037. Figures 10.63c,d show the details of a single butterfly for the encoder and decoder, respectively. Quantization in layer III is nonuniform. The quantizer raises the values to be quantized to 3/4 power before quantization. This provides a more consistent SNR. The decoder reverses this operation by dequantizing a value and raising it to 4/3 power. The quantization is performed by is(i) = nint  xr(i) quant3/4 - 0.0946, (10.16) where xr(i) is the absolute value of the signal of subband i, “quant” is the quantization step size, “nint” is the nearest-integer function, and is(i) is the quantized value. As in layers I and II, the quantization is midtread, i.e., values around 0 are quantized to 0 and the quantizer is symmetric about 0. In layers I and II, each subband can have its own scale factor. Layer III uses bands of scale factors. These bands cover several MDCT coefficients each, and their widths are close to the widths of the critical bands. There is a noise allocation algorithm that selects values for the scale factors. Layer III uses Huffman codes to compress the quantized values even further. The encoder produces 18 MDCT coefficients per subband. It sorts the resulting 576 coefficients (= 18Χ32) by increasing order of frequency (for short blocks, there are three sets of coefficients within each frequency). Notice that the 576 MDCT coefficients correspond to 1152 transformed audio samples. The set of sorted coefficients is divided into 1052 10. Audio Compression for(sb=1; sb<32; sb++) for(i=0; i<8; i++){ xar[18*sb-1-i]=xr[18*sb-1-i]cs[i]-xr[18*sb+i]ca[i] xar[18*sb+i]=xr[18*sb+i]cs[i]+xr[18*sb-1-i]ca[i] } previous block current block (i) i=0,1,...,7 csi csi cai -cai previous block current block (i) i=0,1,...,7 csi csi cai -cai (c) (d) (b) (a) (0) (1) (2) Five more butterflies (0) (1) (2) Five more butterflies X575 X558 X557 X540 X539 18 18 (0) (1) (2) Five more butterflies (0) (1) (2) Five more butterflies X36 X35 X18 X17 X0 18 18 Figure 10.63: Layer III Aliasing Reduction Butterfly. 10.14 MPEG-1/2 Audio Layers 1053 three regions, and each region is encoded with a different set of Huffman codes. This is because the values in each region have a different statistical distribution. The values for the higher frequencies tend to be small and have runs of zeros, whereas the values for the lower frequencies tend to be large. The code tables are provided by the standard (32 tables on pages 54–61 of [ISO/IEC 93]). Dividing the quantized values into three regions also helps to control error propagation. Starting at the values for the highest frequencies, where there are many zeros, the encoder selects the first region as the continuous run of zeros from the highest frequency. It is rare, but possible, not to have any run of zeros. The run is limited to an even number of zeros. This run does not have to be encoded, since its value can be deduced from the sizes of the other two regions. Its size, however, should be even, since the other two regions code their values in even-numbered groupings. The second region consists of a continuous run of the three values -1, 0, and 1. This is called the “count1” region. Each Huffman code for this region encodes four consecutive values, so the number of codes must be 34 = 81. The length of this region must, of course, be a multiple of 4. The third region, known as the “big values” region, consists of the remaining values. It is (optionally) further divided into three subregions, each with its own Huffman code table. Each Huffman code encodes two values. The largest Huffman code table specified by the standard has 16Χ16 codes. Larger values are encoded using an escape mechanism. A frame F in layer III is organized as follows: It starts with the usual 32-bit header, which is followed by the optional 16-bit CRC. This is followed by 136 bits (for single channel) and 256 bits (for dual channel) of side information. The size of side information for MPEG/Audio Layer III, for single and dual channel, are listed in Table 10.64. The last part of the frame is the main data. The side information is followed by a segment of main data (the side information contains, among other data, the length of the segment) but the data in this segment does not have to be that of frame F ! The segment may contain main data from several frames because of the encoder’s use of a bit reservoir. The concept of the bit reservoir has already been mentioned. The encoder can borrow bits from this reservoir when it decides to increase the number of quantization levels because it suspects preechoes. The encoder can also donate bits to the reservoir when it needs fewer than the average number of bits to encode a frame. Borrowing, however, can be done only from past donations; the reservoir cannot have a negative number of bits. The side information of a frame includes a 9-bit pointer that points to the start of the main data for the frame, and the entire concept of main data, fixed-size segments, pointers, and bit reservoirs is illustrated in Figure 10.65. In this diagram, frame 1 needed only about half of its bit allocation, so it left the other half in the reservoir, where it was eventually used by frame 2. Frame 2 needed a little additional space in its “own” segment, leaving the rest of the segment in the reservoir. This was eventually used by frames 3 and 4. Frame 3 did not need any of its own segment, so the entire segment was left in the reservoir, and was eventually used by frame 4. That frame also needed some of its own segment, and the rest was used by frame 5. Bit allocation in layer III is similar to that in the other layers, but includes the added complexity of noise allocation. The encoder (Figure 10.66) computes the bit allocation, 1054 10. Audio Compression Field name Single channel Dual channel main Data 9 9 private bits= 5 3 scfsi: 4 8 part2 3length = 12 Χ2 12Χ 4 big values = 9 Χ2 9Χ 4 global gain = 8 Χ2 8Χ 4 scalefac compress = 4 Χ2 4Χ 4 preflag: 1 Χ2 1Χ 4 scalefac scale: 1 Χ2 1Χ 4 count1table Select = 1 Χ2 1Χ 4 window Switching flag = 1 Χ2 1Χ 4 Additional info = 22 Χ2 22Χ 4 (dependent on window switching flag) Total (bits) 136 256 Table 10.64: Size of Side Info for MPEG Audio Layer III. Bits left in reservoir, eventually 1 2 2 3 4 4 4 5 header & side info frame 2 header & side info frame 1 header & side info frame 3 header & side info frame 4 header & side info frame 5 to be used by another frame Figure 10.65: Layer III Compressed Stream. performs the actual quantization of the subband signals, encodes them with the Huffman codes, and counts the total number of bits generated in this process. This is the bit allocation inner loop. The noise allocation algorithm (also called analysis-by-synthesis procedure) becomes an outer loop where the encoder calculates the quantization noise (i.e., it dequantizes and reconstructs the subband signals and computes the differences between each signal and its reconstructed counterpart). If it finds that certain scalefactor bands have more noise than what the psychoacoustic model allows, the encoder increases the number of quantization levels for these bands, and repeats the process. This process terminates when any of the following three conditions becomes true: 1. All the scalefactor bands have the allowed noise or less. 2. The next iteration would require a requantization of ALL the scalefactor bands. 3. The next iteration would need more bits than are available in the bit reservoir.  Exercise 10.13: Layer III is extremely complex and hard to implement. In view of this, how is it that there are so many free and low-cost programs available to play .mp3 audio files on all computer platforms? 10.15 Advanced Audio Coding (AAC) 1055 Start Start Start Quantization Maximum of all quantized values within table range? Calculate run length of values less or equal one at the upper end of the spectrum Bit count for the coding of the values less or equal one at the upper end of the spectrum Divide the rest of the spectral values into 2 or 3 subregions Choose code table for each subregion Bit count for each subregion Overall bit sum less than available bit? Increase quantizer step size Increase quantizer step size Calculate the distortion for each scalefactor band Save scalefactors Preemphasis Amplify scalefactor bands with more than the allowed distortion All scalefactor bands amplified? Amplification of all bands below upper limit? At least one band with more than the allowed distortion? Restore scaling factors Calculation of available bits Reset of iteration variables All spectral values zero? Outer iteration loop Inner iteration loop Calculate the number of unused bits Return yes yes yes no no no Return Return Layer III Iteration Loop Layer III Outer Iteration Loop Layer III Inner Iteration Loop Figure 10.66: Layer III Iteration Loop. 10.15 Advanced Audio Coding (AAC) The development, by Philips and Sony, of the familiar compact disc (CD), started around 1974. In June 1980, the two companies agreed on a common CD standard and in 1981 this standard was approved by the Digital Audio Disc committee. (Notice the spelling disc, as opposed to a magnetic disk.) The standard, which has been used ever since, includes hardware and software specifications of the signal format, disc material and dimensions, error-correcting code, and many other features. What strikes us as unexpected and unusual, though, is that our music CDs record the audio in uncompressed format. The raw audio samples, typically 44,100 16-bit samples per second of audio, are written on the disc with a sophisticated error-correcting code, but without any attempt to compress them. Current CDs have a capacity of about 80 minutes of sound. This can accommodate many songs, and is enough for a concerto or two and for most symphonies (although 1056 10. Audio Compression Mahler’s third symphony, at about 100 minutes, is a notable exception that comes to mind). An opera, however, requires two or more CDs (Wagner’s Ring Cycle, which clocks in at 18 hours, requires 14–16 CDs), as also do the collected works of many composers. Thus, compression can be quite useful. Currently, the ever popular mp3 format routinely achieves compression factors of 18–20 and can reduce a music library of 100 CDs to just five or six CDs and eventually to just a single DVD. Considering all this, it is natural to wonder about the lack of audio compression in the original CD standard. The best explanation is that engineers simply did not believe that audio can be compressed lossily and efficiently, to perhaps 20% of its original size, without a noticeable loss of quality. A common term used by audio experts in the 1970s and 80s was “golden ears,” which refers to audiophiles who claim they can distinguish between live music and recorded music, especially music played from a compressed format. Thus, when the Moving Pictures Experts Group (MPEG) came into being, it concentrated on video compression. It was only in 1988 that several audio pioneers formed the audio group within MPEG (see also page 1031) and started the process that brought us the three MPEG-1 and MPEG-2 layers, followed by the advanced audio compression (AAC) standard. The AAC compression standard was originally part of the MPEG-2 project, and was later augmented and extended as part of MPEG-4. Some important references for this efficient and complex method are [Bosi and Goldberg 03], [Brandenburg 99], [Pereira and Ebrahimi 02], and [ISO/IEC 03] (the latter is the formal specification of the standard and is not very readable). AAC is the result of an intensive international effort involving companies and individuals. The project started in November 1994, when a number of proposals were submitted to the MPEG-2 committee. The committee came up with a preliminary structure of AAC as a set of modules that constitute independent parts of the overall system. Each module was developed, implemented, and tested separately before becoming part of the final product. The main modules are the following: Gain control Filterbank Prediction Quantization and coding Noiseless Huffman coding Bitstream multiplexing Temporal noise shaping (TNS) Mid/side (M/S) stereo coding Intensity stereo coding These modules are listed in Figures 10.72 and 10.73 and most of them are described in this section. Today, virtually all producers and consumers of audio agree that audio compression (both lossy and lossless) is effective and is very useful (notice the popularity of mp3 players). Even more, those familiar with the MPEG committees and the way they 10.15 Advanced Audio Coding (AAC) 1057 operate agree that the MPEG process of advertising for algorithms, selecting promising candidates, and implementing and testing them, has resulted in the best audio compression methods, while also encouraging compatibility among many different types of audio equipment. Another aspect of the success of MPEG is that the main methods currently used for audio compression came from this organization and not from commercial developers who try to lock users into proprietary algorithms and software. The chief reason for the success of MPEG audio is its development process. The entire process of selecting, testing, and adopting a compression standard is open and competitive. The MPEG audio committee includes academic researchers and industry representatives. They meet periodically to set goals for the next method, to advertise those goals, to receive, study, and discuss ideas from many contributors, and to implement and test the most promising ideas. The first success of the MPEG audio group came in 1991 with MPEG-1. This early video compression standard already incorporated the three layers discussed in Section 10.14. MPEG-2 was finalized in 1997. It added a few minor enhancements to the three layers, but its main contribution to audio compression was the early version of the advanced audio compression (AAC) standard [ISO/IEC 03]. [Layer 1 of MPEG-1/2 is rarely used. Layer 2 is used for DAB (digital audio broadcasting) in Europe, in audio for video, and in broadcast delivery systems. Layer 3 is, of course, widely used—under the name mp3, derived from its file extension—in consumer products as well as in broadcast codecs.] Thus, the full name of mp3 is MPEG-1 and MPEG-2 layer 3. A somewhat shorter name is MPEG-1/2 layer 3. It has long been observed by users and readers that the term “layers” is confusing and unfortunate. Perhaps “levels” or “versions” would have been a better choice. There is also often a confusion of MPEG-2 with level 2. MPEG-3 was intended for HDTV compression but was found to be redundant (it was very similar to MPEG-2) and was merged with MPEG-2 (Section 9.6). The next MPEG design, MPEG-4, was completed in 1999 (Section 9.7). It incorporated several enhancements to AAC and added the AAC-LD (low delay, Section 10.15.3) module. Work on MPEG-7 is currently underway, but many curious souls ask the natural question, What about MPEGs 5 and 6? The official MPEG literature says nothing about this jump in numbering, but it seems (at least this is what one participant in the MPEG meetings recalls) that someone proposed to continue the integer sequence 1, 2, and 4 with 8 (because these are the first powers of 2 and are written in binary as 10...02). In response, someone else proposed (perhaps as a joke) to select 7 instead of 8 because MPEG-7 is supposed to be so different from its predecessors and because 7 is a lucky number. Why is 7 considered a lucky number? No one can say for sure, but here are seven examples that justify this title: The Pythagoreans called 7 the perfect number. They called 3 and 4 (the triangle and the square) the perfect figures. 1058 10. Audio Compression The ancients knew seven planets, the sun, the moon, Mercury, Venus, Mars, Jupiter and Saturn. Similarly, there were seven wonders in the ancient world. In ancient writings and lore, the number 7 plays an important role. The Bible mentions this number many times (the 7 days of the week, the 7 sabbatical years, the 7 years of famine, the 7 years of plenty, the 7 years occupied in the building of King Solomon’s Temple). The Arabians had seven holy temples. In Persian mysteries there were seven spacious caverns through which the aspirants had to pass. The Goths had seven deities, as did the Romans, from whose names are derived our days of the week. This preoccupation with 7 has found its way to modern times in the form of Snow White and the seven dwarfs, and other works of literature. The seven deadly sins (introduced by St. Gregory The Great around 600) are pride, envy, anger, avarice, sadness, gluttony, and lust. The Trumpton fire brigade (Pugh, Pugh, Barney, Magrew, Cuthburt, Dibble and Grub). Trumpton was a stop-motion children’s television show in 1967. Up until the 15th century, the world was believed to have seven seas. The Red sea, the Mediterranean, the Persian gulf, the Black sea, the Adriatic sea, the Caspian sea, and the Indian ocean. In Chinese culture, the 7th day of the first moon of the lunar year is known as Human’s Day and is celebrated as the universal birthday of all humans. The scientists, engineers, and experts who contributed to the three layers of MPEG- 1 and MPEG-2 have hit on a practical and efficient approach to audio compression, the so-called perceptual coding. The idea is to identify those parts of an audio stream that are not fully perceived by the ear/brain system and quantize them or even delete them completely. What is actually quantized is frequency coefficients (often referred to as spectral coefficients) and not the audio samples themselves. It is this principle of acoustic masking that made first mp3 and later, AAC, possible. The human ear (Section 10.3) is very sensitive, but it is not a precision instrument. Its sensitivity depends heavily on the frequency of the sound (as well as on other factors such as age and environment). Figure 10.67 shows the normal threshold of the ear (a curve determined by many sensitive experiments) and how it depends on the frequency (see also Figure 10.5a). Notice that the frequency scale is logarithmic because of its wide interval (from 20 Hz to 20 kHz). Any sound below this threshold is inaudible, and the figure shows that the threshold is high at both low and high frequencies. In order for a sound at these frequencies to be audible, it has to be loud; its amplitude has to exceed the threshold. The threshold is lowest at the frequency range of 3–5 kHz, indicating that the ear is very sensitive at this range and can detect very faint sounds. The point is that the normal threshold can be momentarily perturbed by loud sounds. The figure shows how a loud sound at a frequency of 300 Hz raises the normal threshold for frequencies from 80 Hz to about 1 kHz. The result is that the sound labeled “x,” at 150 Hz, which is above the normal threshold, is now located below the new, perturbed threshold and is therefore masked; it cannot be heard and its audio samples can be deleted (see also Figure 10.5b). 10.15 Advanced Audio Coding (AAC) 1059 20 50 100 200 500 1000 2000 5000 10000 20000 20 40 60 0 Masking sound Masked sound Inaudible sound Inaudible sound dB x Hz Normal threshold Masking threshold Figure 10.67: Audio Frequency Masking. Time is also a factor in acoustic masking. Any perturbation of the normal threshold is temporary and normally decays in less than a second. Figure 10.68 shows a loud, 60- dB sound (at frequency f) that lasts 5 ms. A new threshold starts at 60 dB and decays almost completely in about 500 ms. Any sounds at or around frequency f that happen to be below this threshold are inaudible. Sound “x” at 30 dB is inaudible because it occurs only 5 ms after the masking sound, but the same 30-dB sound “y” occurring 10 ms later is fully audible. The ear becomes overwhelmed by the loud sound to such an extent that it cannot perceive other sounds for a while. dB 5 5 ms 0 20 40 60 10 20 50 100 200 500 Masking sound x y Inaudible sound Figure 10.68: Audio Masking in Time. Figure 10.69 illustrates the concept of acoustic masking in both frequency and time. A masking sound (in the form of a rectangle) creates a surface that is spread over neighboring frequencies and over time. Any sounds with amplitudes and frequencies under this surface are masked. 1060 10. Audio Compression Amplitude (dB) Frequency Time Masking sound Figure 10.69: Audio Masking In Frequency and Time. Filter bank Perceptual model Audio Quantization Coding Output samples stream Figure 10.70: Block Diagram of a Perceptual Encoder. Based on this behavior of the ear, a time/frequency (T/F) lossy audio compression algorithm works in the following steps (Figure 10.70): The encoder goes over the audio samples in the input stream and groups them in overlapping ranges (windows) of samples. Each window is then filtered to convert its audio samples to frequency coefficients. Each coefficient corresponds to a small interval of frequencies (a frequency band) and indicates the intensity of the sound in the band. Use the frequency coefficients to identify the masking sounds. A perceptual (or psychoacoustic) model then estimates the height of the mask curve at each frequency band. This height is determined by the normal threshold and by perturbations to it, caused by masking sounds at various frequencies and in various times in the past. When the updated threshold curve is ready, it is used to determine which frequency coefficients are unimportant. These coefficients are now quantized. The quantized coefficients are encoded by a variable-length code. The codes are written on the output stream together with other side information needed by the decoder. An important point is that audio masking depends on the frequency of sound, but the data to be compressed is in the form of audio samples. Thus, the first task of the encoder is to take a set of audio samples and compute the frequencies of the sound wave expressed by them. (In practice, intervals of frequencies, known as frequency bands, are identified and one frequency (or spectral) coefficient is computed per band. The bands must not exceed the width of the ear’s critical bands discussed in Section 10.3.) This is done by means of a bank of polyphase filters. The way this is done in layer 3 10.15 Advanced Audio Coding (AAC) 1061 (mp3) is discussed in Section 10.14.1 and particularly in Figure 10.48. In mp3, the most-recent 512 audio samples are stored in a buffer and are used to compute spectral coefficients for 32 frequency bands. Each spectral coefficient indicates the amplitude of the audio (specifically, that part of the audio described by the 512 samples) in one frequency band. The next 32 audio samples are then shifted into the buffer and the process is repeated. AAC performs a similar computation, but uses a buffer with 2048 audio samples, computes 1024 spectral coefficients, each indicating the amplitude of the audio in a 23.4-Hz-wide frequency band, and shifts the next 1024 samples into the buffer. [The spectral coefficients are later used by the decoder to reconstruct the audio samples, but these samples are different from the original ones because the spectral coefficients are quantized to achieve compression.] The next component of mp3 is a psychoacoustic model (Section 10.14.5). It determines the height of the masking threshold for every frequency band. However, the model is not specified in detail and the idea is that each implementation selects its own model and the success of a particular implementation depends on the complexity of the model and its resemblance to the actual behavior of the ear/brain system. The AAC standard specification similarly employs a psychoacoustic model without specifying its precise operation. (See flowchart on page 144 of [ISO/IEC 03].) For each frequency band, the encoder compares the spectral coefficient to the threshold of the band. If the spectral coefficient is smaller than the threshold, the coefficient is quantized. This is the main source of compression in the three layers of MPEG-1/2 as well as in AAC, and it is lossy. The amount of quantization depends on how much smaller the coefficient is than the threshold. The amount of quantization is crucial. If a coefficient is quantized too much, the decoder would generate audio samples much different from the original, resulting in perceptible noise and poor audio quality. On the other hand, the result of insufficient (too fine) quantization is low compression that doesn’t achieve the target bitrate specified by the user. AAC employs several tools that improve quantization as follows: Temporal noise shaping (TNS) is a sophisticated algorithm that minimizes the effect of temporal spread. This improves mostly the quantization (and hence the compression) of voice signals. A prediction module improves the performance of the quantizer in cases where the original audio features patterns, such as high tonality (sinusoid-like sound). Perceptual noise shaping (PNS) results in a finer control of quantization, which saves bits and improves compression. The next step is coding. The quantized frequency coefficients are coded. This step improves compression because the coefficients are replaced by variable-length codes, but this added compression is lossless (the AAC literature refers to it as “noiseless”). Both mp3 and AAC use Huffman codes. The former encoding process is described in Section 10.14.6. The latter uses 12 Huffman code tables. To select a code table, the encoder selects a group of two or four consecutive quantized frequency coefficients and selects a code table depending on the group size and on the largest absolute value of the coefficients in the group. Huffman codes for coefficients up to ±16 are provided, but there is also an escape mechanism that allows coding of coefficients of up to ±8,191. 1062 10. Audio Compression The last step is to collect the Huffman codes, add flags, control codes, and other information needed by the decoder, and multiplex all this to create the final output stream. Two unusual features of mp3 are the bit reservoir and the joint stereo mode. The user specifies the quality of the compression by specifying a bitrate. A bitrate of 64 Kb/sec, for example, directs the encoder to compress each second of audio to only 64 K bits. For stereo sound, a sampling rate of 44,100 samples per second results in 88,200 samples (equivalent to 176,400 bytes or 1,411,200 bits) per second. Compressing 1,411,200 bits into 65,536 bits implies a compression factor of about 21.5. This is impressive, but may lead to degredation of the reconstructed sound. Thus, mp3 maintains a bit reservoir. Each time a set of coefficients is quantized, encoded, and output, the encoder figures out how many bits are allowed for the set by the user-specified bitrate and how many bits were actually used. Any “unused” bits are added to the bit reservoir to be used for future sets if needed. If too many bits were used for the compression of the current set, the extra bits are subtracted from the reservoir. If the reservoir doesn’t have enough bits, the encoder repeats its operation with coarser quantization, which saves bits but results in bad reconstructed sound. The joint stereo mode takes advantage of the correlation between the left and right stereo audio channels. Thus, AAC shares the basic features of mp3, with the following improvements: A larger, improved filter bank that computes 1,024 spectral coefficients from each window of 2,048 audio samples, resulting in narrow frequency bands and a frequency resolution much finer than that of mp3. Temporal noise shaping (TNS), a new algorithm that minimizes the effect of temporal masking. This is especially useful for the compression of voice signals. A prediction module improves quantization for periodic audio or audio with patterns. Perceptual noise shaping (PNS) provides fine control of quantization which leads to better compression. In April 2003, Apple Computer brought public attention to AAC by announcing that its iTunes and iPod products would support audio in MPEG-4 AAC format (older iPods require a firmware update). Customers can download music from the iTunes Music Store in a protected version of the format (it uses the file extensions m4a and m4p). This is why AAC has become so associated with Apple hardware and software that many mistakenly believe that AAC stands for “Apple Audio Codec.” MP4 also refers to Mψller-Plesset perturbation theory of the fourth order. 10.15.1 Details of AAC We start with the details of AAC as defined in MPEG-2. Section 10.15.2 discusses the features added to AAC by MPEG-4. Figures 10.72 and 10.73 are block diagrams of the AAC encoder and decoder, respectively. AAC supports three profiles. These are different configurations of the basic algorithm, and they offer different trade-offs between compression efficiency and encoder 10.15 Advanced Audio Coding (AAC) 1063 index profile 0 Main 1 Low Complexity (LC) 2 Scalable Sampling Rate (SSR) 3 Reserved Table 10.71: The Three AAC Profiles. complexity. A 2-bit profile index is written on the compressed stream (Table 10.71) to indicate the profile to the decoder. Index 0. Main profile. In this profile, AAC offers the best compression for any sampling rate and bitrate. All the modules, routines, and tools of the AAC specification (except the gain control tool) are employed. This is the normal profile and it makes sense when enough memory and processing power are available (as is common in current computers). Index 1. Low complexity profile (LC). The gain control tool and prediction are not used. Also, TNS order is limited to 12. Otherwise, this profile is a subset of the main profile, which means that a compressed stream generated by LC can be decoded by the main profile. Index 2. Scalable sampling rate (SSR). This profile is designed to become simplified (to scale down in complexity) when the original audio data consists of limited frequencies (reduced audio bandwidth). The gain-control tool is required in SSR and is used only by SSR (but is not used in the lowest of the four PQF subbands). The prediction and coupling channels are not used. Also, the TNS order and bandwidth are limited. A compressed stream generated by LC can be decoded by SSR, but the resulting audio will be limited to (approximately) 5 kHz because the lowest band of the first filterbank is not used. SSR is simpler than the other profiles. Index 3 is reserved for a future profile. Gain Control. This tool is a preprocessor and is used only in the SSR profile. It consists of modules for filtering, gain detection, and gain modification. Gain control starts by filtering the input (audio samples) into four 6-kHz-wide frequency bands. This filtering is done by a bank of polyphase quadrature filters (PQF). The frequency coefficients in each frequency band are examined by the gain detectors, searching for rapid variations in energy (i.e., consecutive coefficients with very different sizes). The gain modifiers use these results to modify the coefficients so as to compress the dynamics of the input audio. In the AAC decoder, gain control performs the same operations in reverse order; it becomes a postprocessor. The modification are reversed, to restore the original dynamics of the audio signal, and the inverse PQF filterbank is used to generate audio samples. The coefficients of the four PQF frequency bands are computed by hi = 1 4 cos (2i + 1)(2n + 5). 16 Q(n), for 0 . n . 95, and 0 . i . 3, where the first 48 Q(n) values are listed in Table 10.74 and the remaining 48 values are given by Q(n) = Q(95 - n) for 48 . n . 95. 1064 10. Audio Compression input time signal AAC gain control block switching filterbank window length decision psychoacoustic model threshold calculation TNS intensity prediction M/S scaling quantization Huffman coding data control coded audio bitstream stream formatter spectral processing quantization and noiseless coding Figure 10.72: AAC Encoder. 10.15 Advanced Audio Coding (AAC) 1065 AAC gain control block switching filterbank TNS intensity prediction M/S rescaling dequantization Huffman decoding dependently switched coupling dependently switched coupling dependently switched coupling coded audio stream bitstream formatter data control spectral processing dequantization noiseless decoding and input time signal Figure 10.73: AAC Decoder. 1066 10. Audio Compression j Q(j) j Q(j) 0 9.7655291007575512E-05 24 -2.2656858741499447E-02 1 1.3809589379038567E-04 25 -6.8031113858963354E-03 2 9.8400749256623534E-05 26 1.5085400948280744E-02 3 -8.6671544782335723E-05 27 3.9750993388272739E-02 4 -4.6217998911921346E-04 28 6.2445363629436743E-02 5 -1.0211814095158174E-03 29 7.7622327748721326E-02 6 -1.6772149340010668E-03 30 7.9968338496132926E-02 7 -2.2533338951411081E-03 31 6.5615493068475583E-02 8 -2.4987888343213967E-03 32 3.3313658300882690E-02 9 -2.1390815966761882E-03 33 -1.4691563058190206E-02 10 -9.5595397454597772E-04 34 -7.2307890475334147E-02 11 1.1172111530118943E-03 35 -1.2993222541703875E-01 12 3.9091309127348584E-03 36 -1.7551641029040532E-01 13 6.9635703420118673E-03 37 -1.9626543957670528E-01 14 9.5595442159478339E-03 38 -1.8073330670215029E-01 15 1.0815766540021360E-02 39 -1.2097653136035738E-01 16 9.8770514991715300E-03 40 -1.4377370758549035E-02 17 6.1562567291327357E-03 41 1.3522730742860303E-01 18 -4.1793946063629710E-04 42 3.1737852699301633E-01 19 -9.2128743097707640E-03 43 5.1590021798482233E-01 20 -1.8830775873369020E-02 44 7.1080020379761377E-01 21 -2.7226498457701823E-02 45 8.8090632488444798E-01 22 -3.2022840857588906E-02 46 1.0068321641150089E+00 23 -3.0996332527754609E-02 47 1.0737914947736096E+00 Table 10.74: The First 48 Q(n) Coefficients. Figure 10.75 shows the main components of the gain control encoder (part a) and decoder (part b). Filtering. The AAC encoder transforms the audio samples into frequency coefficients by means of a modified DCT (MDCT) transform. The process is similar to that employed by the three layers of MPEG-1 and MPEG-2, but with higher resolution and with improvements. At any given time, the encoder transforms a set of consecurive audio samples referred to as a window. There are two types of windows, long and short. A long window consists of 2,048 consecutive samples and is transformed to produce 1,024 frequency coefficients (a long transform). A short window is 256 samples long and results in 128 coefficients (a short transform). Once the coefficients of a window have been computed, the encoder shifts the audio samples in the buffer by half the window size. Thus, the windows overlap by 50% of their size. Long windows produce many coefficients, so each coefficient corresponds to a narrow frequency band. This leads to precise quantization and thus to a more meaningful loss of data. The data lost in quantization corresponds to those parts of the audio that are not perceived by the ear/brain system. Thus, long windows make sense in those parts of the audio that are either tonal or low frequency. On the other hand, atonal or high-frequency audio varies rapidly, which is why such parts of the audio input lend themselves to better compression with short windows. It is up to the AAC encoder to analyze the input, identify its stationary and transient parts, and use this knowledge to decide when and how often to switch filtering windows. In order to smooth the transition between long and short windows, AAC defines three types of long windows as follows: LONG_WINDOW. This is the normal type of long windows, Many consecutive long windows may be used, they overlap by 1,024 audio samples and each produces 1,024 frequency coefficients. 10.15 Advanced Audio Coding (AAC) 1067 samples audio coefficients frequency samples audio control data gain control data gain controlled time nonoverlapped time signal signal gain gain control tool gain control tool PQF gain detector gain detector gain detector gain modifier gain modifier gain modifier gain compensator & overlapping gain compensator & overlapping gain compensator & overlapping overlapping 256 or 32 MDCT 256 or 32 MDCT 256 or 32 MDCT 256 or 32 MDCT 256 or 32 IMDCT 256 or 32 IMDCT 256 or 32 IMDCT 256 or 32 IMDCT spectral reverse spectral reverse spectral reverse spectral reverse IPQF (a) (b) Figure 10.75: AAC Gain Control (a) Encoder and (b) Decoder. LONG_START_WINDOW. When the encoder decides to switch from a long window to a short one, it uses one window of this type. It has an overlap of 1,024 audio samples with the long window that precedes it and of 128 samples with the short window that follows it. It produces 1,024 frequency coefficients. LONG_STOP_WINDOW. This type of long window is used when the encoder decides to switch from a set of short windows (they always come in sets of eight) to a long window. This type of window has an overlap of 128 audio samples with the short window that precedes it and an overlap of 1,024 samples with the long window that follows it. It also produces 1,024 frequency coefficients. The width of the frequency bands depends on the sampling rate (number of audio samples per second per audio channel), the number of channels (two for stereo), and the number of frequency bands. The AAC literature provides specifications and tables for the following 12 sampling rates (in Hz): 96,000, 88,200, 64,000, 48,000, 44,100, 32,000, 24,000, 22,050, 16,000, 12,000, 11,025, and 8,000. Any other sampling rates must use one of the 12 sets of specifications and tables listed in Table 10.76. For a sampling 1068 10. Audio Compression rate of 48,000 Hz and two stereo channels, each channel is sampled at 24,000 Hz. A long window generates 1,024 frequency coefficients, so each coefficient corresponds to a 24,000/1,024 . 23.4 kHz frequency band. Frequency Specs & tables f . 92, 017 96,000 92, 017 > f . 75, 132 88,200 75, 132 > f . 55, 426 64,000 55, 426 > f . 46, 009 48,000 46, 009 > f . 37, 566 44,100 37, 566 > f . 27, 713 32,000 27, 713 > f . 23, 004 24,000 23, 004 > f . 18, 783 22,050 18, 783 > f . 13, 856 16,000 13, 856 > f . 11, 502 12,000 11, 502 > f . 9, 391 11,025 9, 391 > f 8,000 Table 10.76: 12 Sets of Frequency Specifications. Figure 10.77a shows a typical sequence of three long windows, spanning a total of 4,096 audio samples. Part (b) of the figure shows a LONG_START_WINDOW, followed by eight short windows, followed by a LONG_STOP_WINDOW. 0 1 Gain Samples 0 512 1024 1536 2048 2560 3072 3584 4096 0 1 Gain Samples 0 512 1024 1536 2048 2560 3072 3584 4096 1 2 3 4 5 6 7 8 9 10 (a) (b) Figure 10.77: Overlapping Long and Short Windows for AAC Filtering. The MDCT calculated by the encoder is specified by [compare with Equation (10.14)] Xi,k = 2 N-1 n=0 zi,n cos 2. N (n + n0)(k + 1/2), for 0 . k