\documentclass[b5paper]{book} \usepackage{makeidx} \usepackage{amssymb} \usepackage{color} \usepackage{alltt} \usepackage{graphicx} \usepackage{layout} \def\union{\cup} \def\intersect{\cap} \def\getsrandom{\stackrel{\rm R}{\gets}} \def\cross{\times} \def\cat{\hspace{0.5em} \| \hspace{0.5em}} \def\catn{$\|$} \def\divides{\hspace{0.3em} | \hspace{0.3em}} \def\nequiv{\not\equiv} \def\approx{\raisebox{0.2ex}{\mbox{\small $\sim$}}} \def\lcm{{\rm lcm}} \def\gcd{{\rm gcd}} \def\log{{\rm log}} \def\ord{{\rm ord}} \def\abs{{\mathit abs}} \def\rep{{\mathit rep}} \def\mod{{\mathit\ mod\ }} \renewcommand{\pmod}[1]{\ ({\rm mod\ }{#1})} \newcommand{\floor}[1]{\left\lfloor{#1}\right\rfloor} \newcommand{\ceil}[1]{\left\lceil{#1}\right\rceil} \def\Or{{\rm\ or\ }} \def\And{{\rm\ and\ }} \def\iff{\hspace{1em}\Longleftrightarrow\hspace{1em}} \def\implies{\Rightarrow} \def\undefined{{\rm ``undefined"}} \def\Proof{\vspace{1ex}\noindent {\bf Proof:}\hspace{1em}} \let\oldphi\phi \def\phi{\varphi} \def\Pr{{\rm Pr}} \newcommand{\str}[1]{{\mathbf{#1}}} \def\F{{\mathbb F}} \def\N{{\mathbb N}} \def\Z{{\mathbb Z}} \def\R{{\mathbb R}} \def\C{{\mathbb C}} \def\Q{{\mathbb Q}} \definecolor{DGray}{gray}{0.5} \newcommand{\url}[1]{\mbox{$<${#1}$>$}} \newcommand{\emailaddr}[1]{\mbox{$<${#1}$>$}} \def\twiddle{\raisebox{0.3ex}{\mbox{\tiny $\sim$}}} \def\gap{\vspace{0.5ex}} \makeindex \begin{document} \frontmatter \pagestyle{empty} \title{Multiple-Precision Integer Arithmetic, \\ A Case Study Involving the LibTomMath Project \\ - DRAFT - } \author{\mbox{ %\begin{small} \begin{tabular}{c} Tom St Denis \\ Algonquin College \\ \\ Mads Rasmussen \\ Open Communications Security \\ \\ Greg Rose \\ QUALCOMM Australia \\ \end{tabular} %\end{small} } } \maketitle This text in its entirety is copyright \copyright{}2003 by Tom St Denis. It may not be redistributed electronically or otherwise without the sole permission of the author. The text is freely redistributable as long as it is packaged along with the LibTomMath library in a non-commercial project. Contact the author for other redistribution rights. This text corresponds to the v0.17 release of the LibTomMath project. \begin{alltt} Tom St Denis 111 Banning Rd Ottawa, Ontario K2L 1C3 Canada Phone: 1-613-836-3160 Email: tomstdenis@iahu.ca \end{alltt} This text is formatted to the international B5 paper size of 176mm wide by 250mm tall using the \LaTeX{} {\em book} macro package and the Perl {\em booker} package. \tableofcontents \listoffigures \chapter*{Preface} Blah. \mainmatter \pagestyle{headings} \chapter{Introduction} \section{Multiple Precision Arithmetic} \subsection{The Need for Multiple Precision Arithmetic} The most prevalent use for multiple precision arithmetic (\textit{often referred to as bignum math}) is within public key cryptography. Algorithms such as RSA, Diffie-Hellman and Elliptic Curve Cryptography require large integers in order to resist known cryptanalytic attacks. Typical modern programming languages such as C and Java only provide small single-precision data types which are incapable of precisely representing integers which are often hundreds of bits long. For example, consider multiplying $1,234,567$ by $9,876,543$ in C with an ``unsigned long'' data type. With an x86 machine the result is $4,136,875,833$ while the true result is $12,193,254,061,881$. The original inputs were approximately $21$ and $24$ bits respectively. If the C language cannot multiply two relatively small values together precisely how does anyone expect it to multiply two values that are considerably larger? Most advancements in fast multiple precision arithmetic stem from the desire for faster cryptographic primitives. However, cryptography is not the only field of study that can benefit from fast large integer routines. Another auxiliary use for multiple precision integers is high precision floating point data types. The basic IEEE standard floating point type is made up of an integer mantissa $q$ and an exponent $e$. Numbers are given in the form $n = q \cdot b^e$ where $b = 2$ is specified. Since IEEE is meant to be implemented in hardware the precision of the mantissa is often fairly small (\textit{23, 48 and 64 bits}). Since the mantissa is merely an integer a large multiple precision integer could be used. In effect very high precision floating point arithmetic could be performed. This would be useful where scientific applications must minimize the total output error over long simulations. \subsection{Multiple Precision Arithmetic} \index{multiple precision} Multiple precision arithmetic attempts to the solve the shortcomings of single precision data types such as those from the C and Java programming languages. In essence multiple precision arithmetic is a set of operations that can be performed on members of an algebraic group whose precision is not fixed. The algorithms when implemented to be multiple precision can allow a developer to work with any practical precision required. Typically the arithmetic over the ring of integers denoted by $\Z$ is performed by routines that are collectively and casually referred to as ``bignum'' routines. However, it is possible to have rings of polynomials as well typically denoted by $\Z/p\Z \left [ X \right ]$ which could have variable precision (\textit{or degree}). This text will discuss implementation of the former, however implementing polynomial basis routines should be relatively easy after reading this text. \subsection{Benefits of Multiple Precision Arithmetic} \index{precision} \index{accuracy} Precision of the real value to a given precision is defined loosely as the proximity of the real value to a given representation. Accuracy is defined as the reproducibility of the result. For example, the calculation $1/3 = 0.25$ is imprecise but can be accurate provided it is reproducible. The benefit of multiple precision representations over single precision representations is that often no precision is lost while representing the result of an operation which requires excess precision. For example, the multiplication of two $n$-bit integers requires at least $2n$ bits to represent the result. A multiple precision system would augment the precision of the destination to accomodate the result while a single precision system would truncate excess bits to maintain a fixed level of precision. Multiple precision representations allow for the precision to be very high (\textit{if not exacting}) but at a cost of modest computer resources. The only reasonable case where a multiple precision system will lose precision is when emulating a floating point data type. However, with multiple precision integer arithmetic no precision is lost. \subsection{Basis of Operations} At the heart of all multiple precision integer operations are the ``long-hand'' algorithms we all learned as children in grade school. For example, to multiply $1,234$ by $981$ the student is not taught to memorize the times table for $1,234$, instead they are taught how to long-multiply. That is to multiply each column using simple single digit multiplications, line up the partial results, and add the resulting products by column. The representation that most are familiar with is known as decimal or formally as radix-10. A radix-$n$ representation simply means there are $n$ possible values per digit. For example, binary would be a radix-2 representation. In essence computer based multiple precision arithmetic is very much the same. The most notable difference is the usage of a binary friendly radix. That is to use a radix of the form $2^k$ where $k$ is typically the size of a machine register. Also occasionally more optimal algorithms are used to perform certain operations such as multiplication and squaring instead of traditional long-hand algorithms. \section{Purpose of This Text} The purpose of this text is to instruct the reader regarding how to implement multiple precision algorithms. That is to not only explain the core theoretical algorithms but also the various ``house keeping'' tasks that are neglected by authors of other texts on the subject. Texts such as \cite[HAC]{HAC} and \cite{TAOCPV2} give considerably detailed explanations of the theoretical aspects of the algorithms and very little regarding the practical aspects. How an algorithm is explained and how it is actually implemented are two very different realities. For example, algorithm 14.7 on page 594 of HAC lists a relatively simple algorithm for performing multiple precision integer addition. However, what the description lacks is any discussion concerning the fact that the two integer inputs may be of differing magnitudes. Similarly the division routine (\textit{Algorithm 14.20, pp. 598}) does not discuss how to handle sign or handle the dividend's decreasing magnitude in the main loop (\textit{Step \#3}). As well as the numerous practical oversights both of the texts do not discuss several key optimal algorithms required such as ``Comba'' and Karatsuba multipliers and fast modular inversion. These optimal algorithms are vital to achieve any form of useful performance in non-trivial applications. To solve this problem the focus of this text is on the practical aspects of implementing the algorithms that constitute a multiple precision integer package with light discussions on the theoretical aspects. As a case study the ``LibTomMath''\footnote{Available freely at http://math.libtomcrypt.org} package is used to demonstrate algorithms with implementations that have been field tested and work very well. \section{Discussion and Notation} \subsection{Notation} A multiple precision integer of $n$-digits shall be denoted as $x = (x_n ... x_1 x_0)_{ \beta }$ to be the multiple precision notation for the integer $x \equiv \sum_{i=0}^{n} x_i\beta^i$. The elements of the array $x$ are said to be the radix $\beta$ digits of the integer. For example, $x = (1,2,3)_{10}$ would represent the integer $1\cdot 10^2 + 2\cdot10^1 + 3\cdot10^0 = 123$. A ``mp\_int'' shall refer to a composite structure which contains the digits of the integer as well as auxilary data required to manipulate the data. These additional members are discussed in chapter three. For the purposes of this text a ``multiple precision integer'' and a ``mp\_int'' are synonymous. \index{single-precision} \index{double-precision} \index{mp\_digit} \index{mp\_word} For the purposes of this text a single-precision variable must be able to represent integers in the range $0 \le x < 2 \beta$ while a double-precision variable must be able to represent integers in the range $0 \le x < 2 \beta^2$. Within the source code that will be presented the data type \textbf{mp\_digit} will represent a single-precision type while \textbf{mp\_word} will represent a double-precision type. In several algorithms (\textit{notably the Comba routines}) temporary results will be stored in a double-precision arrays. For the purposes of this text $x_j$ will refer to the $j$'th digit of a single-precision array and $\hat x_j$ will refer to the $j$'th digit of a double-precision array. The $\lfloor \mbox{ } \rfloor$ brackets represent a value truncated and rounded down to the nearest integer. The $\lceil \mbox{ } \rceil$ brackets represent a value truncated and rounded up to the nearest integer. Typically when the $/$ division symbol is used the intention is to perform an integer division. For example, $5/2 = 2$ which will often be written as $\lfloor 5/2 \rfloor = 2$ for clarity. When a value is presented as a fraction such as $5 \over 2$ a real value division is implied. \subsection{Work Effort} \index{big-O} To measure the efficiency of various algorithms a modified big-O notation is used. In this system all single precision operations are considered to have the same cost\footnote{Except where explicitly noted.}. That is a single precision addition, multiplication and division are assumed to take the same time to complete. While this is generally not true in practice it will simplify the discussions considerably. Some algorithms have slight advantages over others which is why some constants will not be removed in the notation. For example, a normal multiplication requires $O(n^2)$ work while a squaring requires $O({{n^2 + n}\over 2})$ work. In standard big-O notation these would be said to be equivalent. However, in the context of the this text the magnitude of the inputs will not approach an infinite size. This means the conventional limit notation wisdom does not apply to the cancellation of constants. Throughout the discussions various ``work levels'' will be discussed. These levels are the $O(1)$, $O(n)$, $O(n^2)$, ..., $O(n^k)$ work efforts. For example, operations at the $O(n^k)$ ``level'' are said to be executed more frequently than operations at the $O(n^m)$ ``level'' when $k > m$. Obviously most optimizations will pay off the most at the higher levels since they represent the bulk of the effort required. \section{Exercises} Within the more advanced chapters a section will be set aside to give the reader some challenging exercises. These exercises are not designed to be prize winning problems, but to be thought provoking. Wherever possible the problems are forward minded stating problems that will be answered in subsequent chapters. The reader is encouraged to finish the exercises as they appear to get a better understanding of the subject material. Similar to the exercises of \cite{TAOCPV2} as explained on pp.\textit{ix} these exercises are given a scoring system. However, unlike \cite{TAOCPV2} the problems do not get nearly as hard as often. The scoring of these exercises ranges from one (\textit{the easiest}) to five (\textit{the hardest}). The following table sumarizes the scoring. \vspace{5mm} \begin{tabular}{cl} $\left [ 1 \right ]$ & An easy problem that should only take the reader a manner of \\ & minutes to solve. Usually does not involve much computer time. \\ & \\ $\left [ 2 \right ]$ & An easy problem that involves a marginal amount of computer \\ & time usage. Usually requires a program to be written to \\ & solve the problem. \\ & \\ $\left [ 3 \right ]$ & A moderately hard problem that requires a non-trivial amount \\ & of work. Usually involves trivial research and development of \\ & new theory from the perspective of a student. \\ & \\ $\left [ 4 \right ]$ & A moderately hard problem that involves a non-trivial amount \\ & of work and research. The solution to which will demonstrate \\ & a higher mastery of the subject matter. \\ & \\ $\left [ 5 \right ]$ & A hard problem that involves concepts that are non-trivial. \\ & Solutions to these problems will demonstrate a complete mastery \\ & of the given subject. \\ & \\ \end{tabular} Essentially problems at the first level are meant to be simple questions that the reader can answer quickly without programming a solution or devising new theory. These problems are quick tests to see if the material is understood. Problems at the second level are also designed to be easy but will require a program or algorithm to be implemented to arrive at the answer. Problems at the third level are meant to be a bit more difficult. Often the answer is fairly obvious but arriving at an exacting solution requires some thought and skill. These problems will almost always involve devising a new algorithm or implementing a variation of another algorithm. Problems at the fourth level are meant to be even more difficult as well as involve some research. The reader will most likely not know the answer right away nor will this text provide the exact details of the answer (\textit{or at least not until a subsequent chapter}). Problems at the fifth level are meant to be the hardest problems relative to all the other problems in the chapter. People who can correctly answer fifth level problems have a mastery of the subject matter at hand. Often problems will be tied together. The purpose of this is to start a chain of thought that will be discussed in future chapters. The reader is encouraged to answer the follow-up problems and try to draw the relevence of problems. \chapter{Introduction to LibTomMath} \section{What is LibTomMath?} LibTomMath is a free and open source multiple precision library written in portable ISO C source code. By portable it is meant that the library does not contain any code that is computer platform dependent or otherwise problematic to use on any given platform. The library has been successfully tested under numerous operating systems including Solaris, MacOS, Windows, Linux, PalmOS and on standalone hardware such as the Gameboy Advance. The library is designed to contain enough functionality to be able to develop applications such as public key cryptosystems. \section{Goals of LibTomMath} Even though the library is written entirely in portable ISO C considerable care has been taken to optimize the algorithm implementations within the library. Specifically the code has been written to work well with the GNU C Compiler (\textit{GCC}) on both x86 and ARMv4 processors. Wherever possible highly efficient algorithms (\textit{such as Karatsuba multiplication, sliding window exponentiation and Montgomery reduction}) have been provided to make the library as efficient as possible. Even with the optimal and sometimes specialized algorithms that have been included the Application Programing Interface (\textit{API}) has been kept as simple as possible. Often generic place holder routines will make use of specialized algorithms automatically without the developer's attention. One such example is the generic multiplication algorithm \textbf{mp\_mul()} which will automatically use Karatsuba multiplication if the inputs are of a specific size. Making LibTomMath as efficient as possible is not the only goal of the LibTomMath project. Ideally the library should be source compatible with another popular library which makes it more attractive for developers to use. In this case the MPI library was used as a API template for all the basic functions. The project is also meant to act as a learning tool for students. The logic being that no easy-to-follow ``bignum'' library exists which can be used to teach computer science students how to perform fast and reliable multiple precision arithmetic. To this end the source code has been given quite a few comments and algorithm discussion points. Often routines have more comments than lines of code. \section{Choice of LibTomMath} LibTomMath was chosen as the case study of this text not only because the author of both projects is one and the same but for more worthy reasons. Other libraries such as GMP, MPI, LIP and OpenSSL have multiple precision integer arithmetic routines but would not be ideal for this text for reasons as will be explained in the following sub-sections. \subsection{Code Base} The LibTomMath code base is all portable ISO C source code. This means that there are no platform dependent conditional segments of code littered throughout the source. This clean and uncluttered approach to the library means that a developer can more readily ascertain the true intent of a given section of source code without trying to keep track of what conditional code will be used. The code base of LibTomMath is also well organized. Each function is in its own separate source code file which allows the reader to find a given function very fast. When compiled with GCC for the x86 processor the entire library is a mere 87,760 bytes (\textit{$116,182$ bytes for ARMv4 processors}). This includes every single function LibTomMath provides from basic arithmetic to various number theoretic functions such as modular exponentiation, various reduction algorithms and Jacobi symbol computation. By comparison MPI which has fewer functions than LibTomMath compiled with the same conditions is 45,429 bytes (\textit{$54,536$ for ARMv4}). GMP which has rather large collection of functions with the default configuration on an x86 Athlon is 2,950,688 bytes. Note that while LibTomMath has fewer functions than GMP it has been used as the sole basis for several public key cryptosystems without having to seek additional outside functions to supplement the library. \subsection{API Simplicity} LibTomMath is designed after the MPI library and shares the API design. Quite often programs that use MPI will build with LibTomMath without change. The function names are relatively straight forward as to what they perform. Almost all of the functions except for a few minor exceptions which as will be discussed are for good reasons share the same parameter passing convention. The learning curve is fairly shallow with the API provided which is an extremely valuable benefit for the student and developer alike. The LIP library is an example of a library with an API that is awkward to work with. LIP uses function names that are often ``compressed'' to illegible short hand. LibTomMath does not share this fault. \subsection{Optimizations} While LibTomMath is certainly not the fastest library (\textit{GMP often beats LibTomMath by a factor of two}) it does feature a set of optimal algorithms for tasks ranging from modular reduction to squaring. GMP and LIP also feature such optimizations while MPI only uses baseline algorithms with no optimizations. LibTomMath is almost always an order of magnitude faster than the MPI library at computationally expensive tasks such as modular exponentiation. In the grand scheme of ``bignum'' libraries LibTomMath is faster than the average library and usually slower than the best libraries such as GMP and OpenSSL by a small factor. \subsection{Portability and Stability} LibTomMath will build ``out of the box'' on any platform equipped with a modern version of the GNU C Compiler (\textit{GCC}). This means that without changes the library will build without configuration or setting up any variables. LIP and MPI will build ``out of the box'' as well but have numerous known bugs. Most notably the author of MPI is not working on his library anymore. GMP requires a configuration script to run and will not build out of the box. GMP and LibTomMath are still in active development and are very stable across a variety of platforms. \subsection{Choice} LibTomMath is a relatively compact, well documented, highly optimized and portable library which seems only natural for the case study of this text. Various source files from the LibTomMath project will be included within the text. However, the reader is encouraged to download their own copy of the library to actually be able to work with the library. \chapter{Getting Started} \section{Library Basics} To begin the design of a multiple precision integer library a primitive data type and a series of primitive algorithms must be established. A data type that will hold the information required to maintain a multiple precision integer must be designed. With this basic data type of a series of low level algorithms for initializing, clearing, growing and optimizing multiple precision integers can be developed to form the basis of the entire library of algorithms. \section{What is a Multiple Precision Integer?} Recall that most programming languages (\textit{in particular C}) only have fixed precision data types that on their own cannot be used to represent values larger than their precision alone will allow. The purpose of multiple precision algorithms is to use these fixed precision data types to create multiple precision integers which may represent values that are much larger. As a well known analogy, school children are taught how to form numbers larger than nine by prepending more radix ten digits. In the decimal system the largest value is only $9$ since the digits may only have values from $0$ to $9$. However, by concatenating digits together larger numbers may be represented. Computer based multiple precision arithmetic is essentially the same concept except with a different radix. What most people probably do not think about explicitly are the various other attributes that describe a multiple precision integer. For example, the integer $154_{10}$ has two immediately obvious properties. First, the integer is positive, that is the sign of this particular integer is positive as oppose to negative. Second, the integer has three digits in its representation. There is an additional property that the integer posesses that does not concern pencil-and-paper arithmetic. The third property is how many digits are allowed for the integer. The human analogy of this third property is ensuring there is enough space on the paper to right the integer. Computers must maintain a strict control on memory usage with respect to the digits of a multiple precision integer. These three properties make up what is known as a multiple precision integer or mp\_int for short. \subsection{The mp\_int structure} The mp\_int structure is the ISO C based manifestation of what represents a multiple precision integer. The ISO C standard does not provide for any such data type but it does provide for making composite data types known as structures. The following is the structure definition used within LibTomMath. \index{mp\_int} \begin{verbatim} typedef struct { int used, alloc, sign; mp_digit *dp; } mp_int; \end{verbatim} The mp\_int structure can be broken down as follows. \begin{enumerate} \item The \textbf{used} parameter denotes how many digits of the array \textbf{dp} contain the digits used to represent a given integer. The \textbf{used} count must not exceed the \textbf{alloc} count. \item The array \textbf{dp} holds the digits that represent the given integer. It is padded with $\textbf{alloc} - \textbf{used}$ zero digits. \item The \textbf{alloc} parameter denotes how many digits are available in the array to use by functions before it has to increase in size. When the \textbf{used} count of a result would exceed the \textbf{alloc} count all of the algorithms will automatically increase the size of the array to accommodate the precision of the result. \item The \textbf{sign} parameter denotes the sign as either zero/positive (\textbf{MP\_ZPOS}) or negative (\textbf{MP\_NEG}). \end{enumerate} \section{Argument Passing} A convention of argument passing must be adopted early on in the development of any library. Making the function prototypes consistent will help eliminate many headaches in the future as the library grows to significant complexity. In LibTomMath the multiple precision integer functions accept parameters from left to right as pointers to mp\_int structures. That means that the source operands are placed on the left and the destination on the right. Consider the following examples. \begin{verbatim} mp_mul(&a, &b, &c); /* c = a * b */ mp_add(&a, &b, &a); /* a = a + b */ mp_sqr(&a, &b); /* b = a * a */ \end{verbatim} The left to right order is a fairly natural way to implement the functions since it lets the developer read aloud the functions and make sense of them. For example, the first function would read ``multiply a and b and store in c''. Certain libraries (\textit{LIP by Lenstra for instance}) accept parameters the other way around. That is the destination on the left and arguments on the right. In truth it is entirely a matter of preference. In the case of LibTomMath the convention from the MPI library has been adopted. Another very useful design consideration is whether to allow argument sources to also be a destination. For example, the second example (\textit{mp\_add}) adds $a$ to $b$ and stores in $a$. This is an important feature to implement since it allows the higher up functions to cut down on the number of variables. However, to implement this feature specific care has to be given to ensure the destination is not modified before the source is fully read. \section{Return Values} A well implemented library, no matter what its purpose, should trap as many runtime errors as possible and return them to the caller. By catching runtime errors a library can be guaranteed to prevent undefined behaviour. In a multiple precision library the only errors that can occur occur are related to inappropriate inputs (\textit{division by zero for instance}) or memory allocation errors. In LibTomMath any function that can cause a runtime error will return an error as an \textbf{int} data type with one of the following values. \index{MP\_OKAY} \index{MP\_VAL} \index{MP\_MEM} \begin{center} \begin{tabular}{|l|l|} \hline \textbf{Value} & \textbf{Meaning} \\ \hline \textbf{MP\_OKAY} & The function was successful \\ \hline \textbf{MP\_VAL} & One of the input value(s) was invalid \\ \hline \textbf{MP\_MEM} & The function ran out of heap memory \\ \hline \end{tabular} \end{center} When an error is detected within a function it should free any memory it allocated and return as soon as possible. The goal is to leave the system in the same state the system was when the function was called. Error checking with this style of API is fairly simple. \begin{verbatim} int err; if ((err = mp_add(&a, &b, &c)) != MP_OKAY) { printf("Error: %d\n", err); exit(EXIT_FAILURE); } \end{verbatim} The GMP library uses C style \textit{signals} to flag errors which is of questionable use. Not all errors are fatal and it was not deemed ideal by the author of LibTomMath to force developers to have signal handlers for such cases. \section{Initialization and Clearing} The logical starting point when actually writing multiple precision integer functions is the initialization and clearing of the integers. These two functions will be used by far the most throughout the algorithms whenever temporary integers are required. Given the basic mp\_int structure an initialization routine must first allocate memory to hold the digits of the integer. Often it is optimal to allocate a sufficiently large pre-set number of digits even considering the initial integer will represent zero. If only a single digit were allocated quite a few re-allocations would occur for the majority of inputs. There is a tradeoff between how many default digits to allocate and how many re-allocations are tolerable. If the memory for the digits has been successfully allocated then the rest of the members of the structure must be initialized. Since the initial state is to represent a zero integer the digits allocated must all be zeroed. The \textbf{used} count set to zero and \textbf{sign} set to \textbf{MP\_ZPOS}. \subsection{Initializing an mp\_int} To initialize an mp\_int the mp\_init algorithm shall be used. The purpose of this algorithm is to allocate the memory required and initialize the integer to a default representation of zero. \begin{figure}[here] \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{mp\_init}. \\ \textbf{Input}. An mp\_int $a$ \\ \textbf{Output}. Allocate memory for the digits and set to a zero state. \\ \hline \\ 1. Allocate memory for \textbf{MP\_PREC} digits. \\ 2. If the allocation failed then return(\textit{MP\_MEM}) \\ 3. for $n$ from $0$ to $MP\_PREC - 1$ do \\ \hspace{3mm}3.1 $a_n \leftarrow 0$\\ 4. $a.sign \leftarrow MP\_ZPOS$\\ 5. $a.used \leftarrow 0$\\ 6. $a.alloc \leftarrow MP\_PREC$\\ 7. Return(\textit{MP\_OKAY})\\ \hline \end{tabular} \end{center} \caption{Algorithm mp\_init} \end{figure} \textbf{Algorithm mp\_init.} The \textbf{MP\_PREC} variable is a simple constant used to dictate minimal precision of allocated integers. It is ideally at least equal to $32$ but can be any reasonable power of two. Steps one and two allocate the memory and account for it. If the allocation fails the algorithm returns immediately to signal the failure. Step three will ensure that all the digits are in the default state of zero. Finally steps four through six set the default settings of the \textbf{sign}, \textbf{used} and \textbf{alloc} members of the mp\_int structure. \index{bn\_mp\_init.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_mp\_init.c \vspace{-3mm} \begin{alltt} 016 017 /* init a new bigint */ 018 int 019 mp_init (mp_int * a) 020 \{ 021 /* allocate ram required and clear it */ 022 a->dp = OPT_CAST calloc (sizeof (mp_digit), MP_PREC); 023 if (a->dp == NULL) \{ 024 return MP_MEM; 025 \} 026 027 /* set the used to zero, allocated digits to the default precision 028 * and sign to positive */ 029 a->used = 0; 030 a->alloc = MP_PREC; 031 a->sign = MP_ZPOS; 032 033 return MP_OKAY; 034 \} \end{alltt} \end{small} The \textbf{OPT\_CAST} type cast on line 22 is designed to allow C++ compilers to build the code out of the box. Microsoft C V5.00 is known to cause problems without the cast. Also note that if the memory allocation fails the other members of the mp\_int will be in an undefined state. The code from line 29 to line 31 sets the default state for a mp\_int which is zero, positive and no used digits. \subsection{Clearing an mp\_int} When an mp\_int is no longer required the memory allocated for it can be cleared from the heap with the mp\_clear algorithm. \begin{figure}[here] \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{mp\_clear}. \\ \textbf{Input}. An mp\_int $a$ \\ \textbf{Output}. The memory for $a$ is cleared. \\ \hline \\ 1. If $a$ has been previously freed then return(\textit{MP\_OKAY}). \\ 2. Free the digits of $a$ and mark $a$ as freed. \\ 3. $a.used \leftarrow 0$ \\ 4. $a.alloc \leftarrow 0$ \\ 5. Return(\textit{MP\_OKAY}). \\ \hline \end{tabular} \end{center} \caption{Algorithm mp\_clear} \end{figure} \textbf{Algorithm mp\_clear.} In steps one and two the memory for the digits are only free'd if they had not been previously released before. This is more of concern for the implementation since it is used to prevent ``double-free'' errors. It also helps catch code errors where mp\_ints are used after being cleared. Similarly steps three and four set the \textbf{used} and \textbf{alloc} to known values which would be easy to spot during debugging. For example, if an mp\_int is expected to be non-zero and its \textbf{used} member is observed to be zero (\textit{due to being cleared}) then an obvious bug in the code has been spotted. \index{bn\_mp\_clear.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_mp\_clear.c \vspace{-3mm} \begin{alltt} 016 017 /* clear one (frees) */ 018 void 019 mp_clear (mp_int * a) 020 \{ 021 if (a->dp != NULL) \{ 022 023 /* first zero the digits */ 024 memset (a->dp, 0, sizeof (mp_digit) * a->used); 025 026 /* free ram */ 027 free (a->dp); 028 029 /* reset members to make debugging easier */ 030 a->dp = NULL; 031 a->alloc = a->used = 0; 032 \} 033 \} \end{alltt} \end{small} The \textbf{if} statement on line 21 prevents the heap from being corrupted if a user double-frees an mp\_int. For example, a trivial case of this bug would be as follows. \begin{verbatim} mp_int a; mp_init(&a); mp_clear(&a); mp_clear(&a); \end{verbatim} Without that check the code would try to free the memory allocated for the digits twice which will cause most standard C libraries to cause a fault. Also by setting the pointer to \textbf{NULL} it helps debug code that may inadvertently free the mp\_int before it is truly not needed. The allocated digits are set to zero before being freed on line 24. This is ideal for cryptographic situations where the mp\_int is a secret parameter. The following snippet is an example of using both the init and clear functions. \begin{small} \begin{verbatim} #include #include #include int main(void) { mp_int num; int err; /* init the bignum */ if ((err = mp_init(&num)) != MP_OKAY) { printf("Error: %d\n", err); return EXIT_FAILURE; } /* do work with it ... */ /* clear up */ mp_clear(&num); return EXIT_SUCCESS; } \end{verbatim} \end{small} \section{Other Initialization Routines} It is often helpful to have specialized initialization algorithms to simplify the design of other algorithms. For example, an initialization followed by a copy is a common operation when temporary copies of integers are required. It is quite beneficial to have a series of simple helper functions available. \subsection{Initializing Variable Sized mp\_int Structures} Occasionally the number of digits required will be known in advance of an initialization. In these cases the mp\_init\_size algorithm can be of use. The purpose of this algorithm is similar to mp\_init except that it will allocate \textit{at least} a specified number of digits. This is ideal to prevent re-allocations when the input size is known. \newpage\begin{figure}[here] \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{mp\_init\_size}. \\ \textbf{Input}. An mp\_int $a$ and the requested number of digits $b$\\ \textbf{Output}. $a$ is initialized to hold at least $b$ digits. \\ \hline \\ 1. $u \leftarrow b\mbox{ (mod }MP\_PREC\mbox{)}$ \\ 2. $v \leftarrow b + 2 \cdot MP\_PREC - u$ \\ 3. Allocate $v$ digits. \\ 4. If the allocation failed then return(\textit{MP\_MEM}). \\ 5. for $n$ from $0$ to $v - 1$ do \\ \hspace{3mm}5.1 $a_n \leftarrow 0$ \\ 6. $a.sign \leftarrow MP\_ZPOS$\\ 7. $a.used \leftarrow 0$\\ 8. $a.alloc \leftarrow v$\\ 9. Return(\textit{MP\_OKAY})\\ \hline \end{tabular} \end{center} \caption{Algorithm mp\_init\_size} \end{figure} \textbf{Algorithm mp\_init\_size.} The value of $v$ is calculated to be at least the requested amount of digits $b$ plus additional padding. The padding is calculated to be at least \textbf{MP\_PREC} digits plus enough digits to make the digit count a multiple of \textbf{MP\_PREC}. This padding is used to prevent trivial allocations from becoming a bottleneck in the rest of the algorithms that depend on this. \index{bn\_mp\_init\_size.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_mp\_init\_size.c \vspace{-3mm} \begin{alltt} 016 017 /* init a mp_init and grow it to a given size */ 018 int 019 mp_init_size (mp_int * a, int size) 020 \{ 021 022 /* pad size so there are always extra digits */ 023 size += (MP_PREC * 2) - (size & (MP_PREC - 1)); 024 025 /* alloc mem */ 026 a->dp = OPT_CAST calloc (sizeof (mp_digit), size); 027 if (a->dp == NULL) \{ 028 return MP_MEM; 029 \} 030 a->used = 0; 031 a->alloc = size; 032 a->sign = MP_ZPOS; 033 034 return MP_OKAY; 035 \} \end{alltt} \end{small} Line 23 will ensure that the number of digits actually allocated is padded up to the next multiple of \textbf{MP\_PREC} plus an additional \textbf{MP\_PREC}. This ensures that the number of allocated digit is always greater than the amount requested. As a result it prevents many trivial memory allocations. The value of \textbf{MP\_PREC} is defined in ``tommath.h'' and must be a power of two. \subsection{Creating a Clone} Another common sequence of operations is to make a local temporary copy of an argument. To initialize then copy a mp\_int will be known as creating a clone. This is useful within functions that need to modify an integer argument but do not wish to actually modify the original copy. The mp\_init\_copy algorithm will perform this very task. \begin{figure}[here] \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{mp\_init\_copy}. \\ \textbf{Input}. An mp\_int $a$ and $b$\\ \textbf{Output}. $a$ is initialized to be a copy of $b$. \\ \hline \\ 1. Init $a$. (\textit{mp\_init}) \\ 2. If the init of $a$ was unsuccessful return(\textit{MP\_MEM}) \\ 3. Copy $b$ to $a$. (\textit{mp\_copy}) \\ 4. Return the status of the copy operation. \\ \hline \end{tabular} \end{center} \caption{Algorithm mp\_init\_copy} \end{figure} \textbf{Algorithm mp\_init\_copy.} This algorithm will initialize a mp\_int variable and copy another previously initialized mp\_int variable into it. The algorithm will detect when the initialization fails and returns the error to the calling algorithm. As such this algorithm will perform two operations in one step. \index{bn\_mp\_init\_copy.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_mp\_init\_copy.c \vspace{-3mm} \begin{alltt} 016 017 /* creates "a" then copies b into it */ 018 int 019 mp_init_copy (mp_int * a, mp_int * b) 020 \{ 021 int res; 022 023 if ((res = mp_init (a)) != MP_OKAY) \{ 024 return res; 025 \} 026 return mp_copy (b, a); 027 \} \end{alltt} \end{small} This will initialize \textbf{a} and make it a verbatim copy of the contents of \textbf{b}. Note that \textbf{a} will have its own memory allocated which means that \textbf{b} may be cleared after the call and \textbf{a} will be left intact. \subsection{Multiple Integer Initializations And Clearings} Occasionally a function will require a series of mp\_int data types to be made available. The mp\_init\_multi algorithm is provided to simplify such cases. The purpose of this algorithm is to initialize a variable length array of mp\_int structures at once. As a result algorithms that require multiple integers only has to use one algorithm to initialize all the mp\_int variables. \begin{figure}[here] \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{mp\_init\_multi}. \\ \textbf{Input}. Variable length array of mp\_int variables of length $k$. \\ \textbf{Output}. The array is initialized such that each each mp\_int is ready to use. \\ \hline \\ 1. for $n$ from 0 to $k - 1$ do \\ \hspace{+3mm}1.1. Initialize the $n$'th mp\_int (\textit{mp\_init}) \\ \hspace{+3mm}1.2. If initialization failed then do \\ \hspace{+6mm}1.2.1. for $j$ from $0$ to $n$ do \\ \hspace{+9mm}1.2.1.1. Free the $j$'th mp\_int (\textit{mp\_clear}) \\ \hspace{+6mm}1.2.2. Return(\textit{MP\_MEM}) \\ 2. Return(\textit{MP\_OKAY}) \\ \hline \end{tabular} \end{center} \caption{Algorithm mp\_init\_multi} \end{figure} \textbf{Algorithm mp\_init\_multi.} The algorithm will initialize the array of mp\_int variables one at a time. As soon as an runtime error is detected (\textit{step 1.2}) all of the previously initialized variables are cleared. The goal is an ``all or nothing'' initialization which allows for quick recovery from runtime errors. Similarly to clear a variable length array of mp\_int structures the mp\_clear\_multi algorithm will be used. Consider the following snippet which demonstrates how to use both routines. \begin{small} \begin{verbatim} #include #include #include int main(void) { mp_int num1, num2, num3; int err; if ((err = mp_init_multi(&num1, &num2, &num3, NULL)) !- MP_OKAY) { printf("Error: %d\n", err); return EXIT_FAILURE; } /* at this point num1/num2/num3 are ready */ /* free them */ mp_clear_multi(&num1, &num2, &num3, NULL); return EXIT_SUCCESS; } \end{verbatim} \end{small} Note how both lists are terminated with the \textbf{NULL} variable. This indicates to the algorithms to stop fetching parameters off of the stack. If it is not present the functions will most likely cause a segmentation fault. \index{bn\_mp\_multi.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_mp\_multi.c \vspace{-3mm} \begin{alltt} 016 #include 017 018 int mp_init_multi(mp_int *mp, ...) 019 \{ 020 mp_err res = MP_OKAY; /* Assume ok until proven otherwise */ 021 int n = 0; /* Number of ok inits */ 022 mp_int* cur_arg = mp; 023 va_list args; 024 025 va_start(args, mp); /* init args to next argument from caller */ 026 while (cur_arg != NULL) \{ 027 if (mp_init(cur_arg) != MP_OKAY) \{ 028 /* Oops - error! Back-track and mp_clear what we already 029 succeeded in init-ing, then return error. 030 */ 031 va_list clean_args; 032 033 /* end the current list */ 034 va_end(args); 035 036 /* now start cleaning up */ 037 cur_arg = mp; 038 va_start(clean_args, mp); 039 while (n--) \{ 040 mp_clear(cur_arg); 041 cur_arg = va_arg(clean_args, mp_int*); 042 \} 043 va_end(clean_args); 044 res = MP_MEM; 045 break; 046 \} 047 n++; 048 cur_arg = va_arg(args, mp_int*); 049 \} 050 va_end(args); 051 return res; /* Assumed ok, if error flagged above. */ 052 \} 053 054 void mp_clear_multi(mp_int *mp, ...) 055 \{ 056 mp_int* next_mp = mp; 057 va_list args; 058 va_start(args, mp); 059 while (next_mp != NULL) \{ 060 mp_clear(next_mp); 061 next_mp = va_arg(args, mp_int*); 062 \} 063 va_end(args); 064 \} \end{alltt} \end{small} Both routines are implemented in the same source file since they are typically used in conjunction with each other. \section{Maintenance} A small useful collection of mp\_int maintenance functions will also prove useful. \subsection{Augmenting Integer Precision} When storing a value in an mp\_int sufficient digits must be available to accomodate the entire value without loss of precision. Quite often the size of the array given by the \textbf{alloc} member is large enough to simply increase the \textbf{used} digit count. However, when the size of the array is too small it must be re-sized appropriately to accomodate the result. The mp\_grow algorithm will provide this functionality. \begin{figure}[here] \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{mp\_grow}. \\ \textbf{Input}. An mp\_int $a$ and an integer $b$. \\ \textbf{Output}. $a$ is expanded to accomodate $b$ digits. \\ \hline \\ 1. if $a.alloc \ge b$ then return(\textit{MP\_OKAY}) \\ 2. $u \leftarrow b\mbox{ (mod }MP\_PREC\mbox{)}$ \\ 3. $v \leftarrow b + 2 \cdot MP\_PREC - u$ \\ 4. Re-Allocate the array of digits $a$ to size $v$ \\ 5. If the allocation failed then return(\textit{MP\_MEM}). \\ 6. for n from a.alloc to $v - 1$ do \\ \hspace{+3mm}6.1 $a_n \leftarrow 0$ \\ 7. $a.alloc \leftarrow v$ \\ 8. Return(\textit{MP\_OKAY}) \\ \hline \end{tabular} \end{center} \caption{Algorithm mp\_grow} \end{figure} \textbf{Algorithm mp\_grow.} Step one will prevent a re-allocation from being performed if it was not required. This is useful to prevent mp\_ints from growing excessively in code that erroneously calls mp\_grow. Similar to mp\_init\_size the requested digit count is padded to provide more digits than requested. In step four it is assumed that the reallocation leaves the lower $a.alloc$ digits intact. This is much akin to how the \textit{realloc} function from the standard C library works. Since the newly allocated digits are assumed to contain undefined values they are also initially zeroed. \index{bn\_mp\_grow.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_mp\_grow.c \vspace{-3mm} \begin{alltt} 016 017 /* grow as required */ 018 int 019 mp_grow (mp_int * a, int size) 020 \{ 021 int i; 022 023 /* if the alloc size is smaller alloc more ram */ 024 if (a->alloc < size) \{ 025 /* ensure there are always at least MP_PREC digits extra on top */ 026 size += (MP_PREC * 2) - (size & (MP_PREC - 1)); 027 028 a->dp = OPT_CAST realloc (a->dp, sizeof (mp_digit) * size); 029 if (a->dp == NULL) \{ 030 return MP_MEM; 031 \} 032 033 /* zero excess digits */ 034 i = a->alloc; 035 a->alloc = size; 036 for (; i < a->alloc; i++) \{ 037 a->dp[i] = 0; 038 \} 039 \} 040 return MP_OKAY; 041 \} \end{alltt} \end{small} The first step is to see if we actually need to perform a re-allocation at all. This is tested for on line 24. Similar to mp\_init\_size the same code on line 26 was used to resize the digits requested. A simple for loop from line 34 to line 38 will zero all digits that were above the old \textbf{alloc} limit to make sure the integer is in a known state. \subsection{Clamping Excess Digits} When a function anticipates a result will be $n$ digits it is simpler to assume this is true within the body of the function. For example, a multiplication of a $i$ digit number by a $j$ digit produces a result of at most $i + j$ digits. It is entirely possible that the result is $i + j - 1$ though, with no final carry into the last position. However, suppose the destination had to be first expanded (\textit{via mp\_grow}) to accomodate $i + j - 1$ digits than further expanded to accomodate the final carry. That would be a considerable waste of time since heap operations are relatively slow. The ideal solution is to always assume the result is $i + j$ and fix up the \textbf{used} count after the function terminates. This way a single heap operation (\textit{at most}) is required. However, if the result was not checked there would be an excess high order zero digit. For example, suppose the product of two integers was $x_n = (0x_{n-1}x_{n-2}...x_0)_{\beta}$. The leading zero digit will not contribute to the precision of the result. In fact, through subsequent operations more leading zero digits would accumulate to the point the size of the integer would be prohibitive. As a result even though the precision is very low the representation is excessively large. The mp\_clamp algorithm is designed to solve this very problem. It will trim leading zeros by decrementing the \textbf{used} count until a non-zero leading digit is found. Also in this system, zero is considered to be a positive number which means that if the \textbf{used} count is decremented to zero the sign must be set to \textbf{MP\_ZPOS}. \begin{figure}[here] \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{mp\_clamp}. \\ \textbf{Input}. An mp\_int $a$ \\ \textbf{Output}. Any excess leading zero digits of $a$ are removed \\ \hline \\ 1. while $a.used > 0$ and $a_{a.used - 1} = 0$ do \\ \hspace{+3mm}1.1 $a.used \leftarrow a.used - 1$ \\ 2. if $a.used = 0$ then do \\ \hspace{+3mm}2.1 $a.sign \leftarrow MP\_ZPOS$ \\ \hline \\ \end{tabular} \end{center} \caption{Algorithm mp\_clamp} \end{figure} \textbf{Algorithm mp\_clamp.} As can be expected this algorithm is very simple. The loop on step one is expected to iterate only once or twice at the most. For example, this will happen in cases where there is not a carry to fill the last position. Step two fixes the sign for when all of the digits are zero to ensure that the mp\_int is valid at all times. \index{bn\_mp\_clamp.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_mp\_clamp.c \vspace{-3mm} \begin{alltt} 016 017 /* trim unused digits 018 * 019 * This is used to ensure that leading zero digits are 020 * trimed and the leading "used" digit will be non-zero 021 * Typically very fast. Also fixes the sign if there 022 * are no more leading digits 023 */ 024 void 025 mp_clamp (mp_int * a) 026 \{ 027 while (a->used > 0 && a->dp[a->used - 1] == 0) \{ 028 --(a->used); 029 \} 030 if (a->used == 0) \{ 031 a->sign = MP_ZPOS; 032 \} 033 \} \end{alltt} \end{small} Note on line 27 how to test for the \textbf{used} count is made on the left of the \&\& operator. In the C programming language the terms to \&\& are evaluated left to right with a boolean short-circuit if any condition fails. This is important since if the \textbf{used} is zero the test on the right would fetch below the array. That is obviously undesirable. The parenthesis on line 28 is used to make sure the \textbf{used} count is decremented and not the pointer ``a''. \section*{Exercises} \begin{tabular}{cl} $\left [ 1 \right ]$ & Discuss the relevance of the \textbf{used} member of the mp\_int structure. \\ & \\ $\left [ 1 \right ]$ & Discuss the consequences of not using padding when performing allocations. \\ & \\ $\left [ 2 \right ]$ & Estimate an ideal value for \textbf{MP\_PREC} when performing 1024-bit RSA \\ & encryption when $\beta = 2^{28}$. \\ & \\ $\left [ 1 \right ]$ & Discuss the relevance of the algorithm mp\_clamp. What does it prevent? \\ & \\ $\left [ 1 \right ]$ & Give an example of when the algorithm mp\_init\_copy might be useful. \\ & \\ \end{tabular} \chapter{Basic Operations} \section{Copying an Integer} After the various house-keeping routines are in place, simple algorithms can be designed to take advantage of them. Being able to make a verbatim copy of an integer is a very useful function to have. To copy an integer the mp\_copy algorithm will be used. \newpage\begin{figure}[here] \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{mp\_copy}. \\ \textbf{Input}. An mp\_int $a$ and $b$. \\ \textbf{Output}. Store a copy of $a$ in $b$. \\ \hline \\ 1. Check if $a$ and $b$ point to the same location in memory. \\ 2. If true then return(\textit{MP\_OKAY}). \\ 3. If $b.alloc < a.used$ then grow $b$ to $a.used$ digits. (\textit{mp\_grow}) \\ 4. If failed to grow then return(\textit{MP\_MEM}). \\ 5. for $n$ from 0 to $a.used - 1$ do \\ \hspace{3mm}5.1 $b_{n} \leftarrow a_{n}$ \\ 6. if $a.used < b.used - 1$ then \\ \hspace{3mm}6.1. for $n$ from $a.used$ to $b.used - 1$ do \\ \hspace{6mm}6.1.1 $b_{n} \leftarrow 0$ \\ 7. $b.used \leftarrow a.used$ \\ 8. $b.sign \leftarrow a.sign$ \\ 9. return(\textit{MP\_OKAY}) \\ \hline \end{tabular} \end{center} \caption{Algorithm mp\_copy} \end{figure} \textbf{Algorithm mp\_copy.} Step 1 and 2 make sure that the two mp\_ints are unique. This allows the user to call the copy function with potentially the same input and not waste time. Step 3 and 4 ensure that the destination is large enough to hold a copy of the input $a$. Note that the \textbf{used} member of $b$ may be smaller than the \textbf{used} member of $a$ but a memory re-allocation is only required if the \textbf{alloc} member of $b$ is smaller. This prevents trivial memory reallocations. Step 5 copies the digits from $a$ to $b$ while step 6 ensures that if initially $\vert b \vert > \vert a \vert$, the more significant digits of $b$ will be zeroed. Finally steps 7 and 8 copies the \textbf{used} and \textbf{sign} members over which completes the copy operation. \index{bn\_mp\_copy.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_mp\_copy.c \vspace{-3mm} \begin{alltt} 016 017 /* copy, b = a */ 018 int 019 mp_copy (mp_int * a, mp_int * b) 020 \{ 021 int res, n; 022 023 /* if dst == src do nothing */ 024 if (a == b) \{ 025 return MP_OKAY; 026 \} 027 028 /* grow dest */ 029 if ((res = mp_grow (b, a->used)) != MP_OKAY) \{ 030 return res; 031 \} 032 033 /* zero b and copy the parameters over */ 034 \{ 035 register mp_digit *tmpa, *tmpb; 036 037 /* pointer aliases */ 038 tmpa = a->dp; 039 tmpb = b->dp; 040 041 /* copy all the digits */ 042 for (n = 0; n < a->used; n++) \{ 043 *tmpb++ = *tmpa++; 044 \} 045 046 /* clear high digits */ 047 for (; n < b->used; n++) \{ 048 *tmpb++ = 0; 049 \} 050 \} 051 b->used = a->used; 052 b->sign = a->sign; 053 return MP_OKAY; 054 \} \end{alltt} \end{small} Source lines 23-31 do the initial house keeping. That is to see if the input is unique and if so to make sure there is enough room. If not enough space is available it returns the error and leaves the destination variable intact. The inner loop of the copy operation is contained between lines 34 and 50. Many LibTomMath routines are designed with this source code style in mind, making aliases to shorten lengthy pointers (\textit{see line 38 and 39}) for rapid use. Also the use of nested braces creates a simple way to denote various portions of code that reside on various work levels. Here, the copy loop is at the $O(n)$ level. \section{Zeroing an Integer} Reseting an mp\_int to the default state is a common step in many algorithms. The mp\_zero algorithm will be the algorithm used to perform this task. \begin{figure}[here] \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{mp\_zero}. \\ \textbf{Input}. An mp\_int $a$ \\ \textbf{Output}. Zero the contents of $a$ \\ \hline \\ 1. $a.used \leftarrow 0$ \\ 2. $a.sign \leftarrow$ MP\_ZPOS \\ 3. for $n$ from 0 to $a.alloc - 1$ do \\ \hspace{3mm}3.1 $a_n \leftarrow 0$ \\ \hline \end{tabular} \end{center} \caption{Algorithm mp\_zero} \end{figure} \textbf{Algorithm mp\_zero.} This algorithm simply resets a mp\_int to the default state. \index{bn\_mp\_zero.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_mp\_zero.c \vspace{-3mm} \begin{alltt} 016 017 /* set to zero */ 018 void 019 mp_zero (mp_int * a) 020 \{ 021 a->sign = MP_ZPOS; 022 a->used = 0; 023 memset (a->dp, 0, sizeof (mp_digit) * a->alloc); 024 \} \end{alltt} \end{small} After the function is completed, all of the digits are zeroed, the \textbf{used} count is zeroed and the \textbf{sign} variable is set to \textbf{MP\_ZPOS}. \section{Sign Manipulation} \subsection{Absolute Value} With the mp\_int representation of an integer, calculating the absolute value is trivial. The mp\_abs algorithm will compute the absolute value of an mp\_int. \begin{figure}[here] \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{mp\_abs}. \\ \textbf{Input}. An mp\_int $a$ \\ \textbf{Output}. Computes $b = \vert a \vert$ \\ \hline \\ 1. Copy $a$ to $b$. (\textit{mp\_copy}) \\ 2. If the copy failed return(\textit{MP\_MEM}). \\ 3. $b.sign \leftarrow MP\_ZPOS$ \\ 4. Return(\textit{MP\_OKAY}) \\ \hline \end{tabular} \end{center} \caption{Algorithm mp\_abs} \end{figure} \textbf{Algorithm mp\_abs.} This algorithm computes the absolute of an mp\_int input. As can be expected the algorithm is very trivial. \index{bn\_mp\_abs.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_mp\_abs.c \vspace{-3mm} \begin{alltt} 016 017 /* b = |a| 018 * 019 * Simple function copies the input and fixes the sign to positive 020 */ 021 int 022 mp_abs (mp_int * a, mp_int * b) 023 \{ 024 int res; 025 if ((res = mp_copy (a, b)) != MP_OKAY) \{ 026 return res; 027 \} 028 b->sign = MP_ZPOS; 029 return MP_OKAY; 030 \} \end{alltt} \end{small} \subsection{Integer Negation} With the mp\_int representation of an integer, calculating the negation is also trivial. The mp\_neg algorithm will compute the negative of an mp\_int input. \newpage\begin{figure}[here] \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{mp\_neg}. \\ \textbf{Input}. An mp\_int $a$ \\ \textbf{Output}. Computes $b = -a$ \\ \hline \\ 1. Copy $a$ to $b$. (\textit{mp\_copy}) \\ 2. If the copy failed return(\textit{MP\_MEM}). \\ 3. If $a.sign = MP\_ZPOS$ then do \\ \hspace{3mm}3.1 $b.sign = MP\_NEG$. \\ 4. else do \\ \hspace{3mm}4.1 $b.sign = MP\_ZPOS$. \\ 5. Return(\textit{MP\_OKAY}) \\ \hline \end{tabular} \end{center} \caption{Algorithm mp\_neg} \end{figure} \textbf{Algorithm mp\_neg.} This algorithm computes the negation of an input. \index{bn\_mp\_neg.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_mp\_neg.c \vspace{-3mm} \begin{alltt} 016 017 /* b = -a */ 018 int 019 mp_neg (mp_int * a, mp_int * b) 020 \{ 021 int res; 022 if ((res = mp_copy (a, b)) != MP_OKAY) \{ 023 return res; 024 \} 025 b->sign = (a->sign == MP_ZPOS) ? MP_NEG : MP_ZPOS; 026 return MP_OKAY; 027 \} \end{alltt} \end{small} \section{Small Constants} \subsection{Setting Small Constants} Often a mp\_int must be set to a relatively small value such as $1$ or $2$. For these cases the mp\_set algorithm is useful. \newpage\begin{figure} \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{mp\_set}. \\ \textbf{Input}. An mp\_int $a$ and a digit $b$ \\ \textbf{Output}. Make $a$ equivalent to $b$ \\ \hline \\ 1. Zero $a$ (\textit{mp\_zero}). \\ 2. $a_0 \leftarrow b \mbox{ (mod }\beta\mbox{)}$ \\ 3. $a.used \leftarrow \left \lbrace \begin{array}{ll} 1 & \mbox{if }a_0 > 0 \\ 0 & \mbox{if }a_0 = 0 \end{array} \right .$ \\ \hline \end{tabular} \end{center} \caption{Algorithm mp\_set} \end{figure} \textbf{Algorithm mp\_set.} This algorithm sets a mp\_int to a small single digit value. Step number 1 ensures that the integer is reset to the default state. The single digit is set (\textit{modulo $\beta$}) and the \textbf{used} count is adjusted accordingly. \index{bn\_mp\_set.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_mp\_set.c \vspace{-3mm} \begin{alltt} 016 017 /* set to a digit */ 018 void 019 mp_set (mp_int * a, mp_digit b) 020 \{ 021 mp_zero (a); 022 a->dp[0] = b & MP_MASK; 023 a->used = (a->dp[0] != 0) ? 1 : 0; 024 \} \end{alltt} \end{small} Line 21 calls mp\_zero() to clear the mp\_int and reset the sign. Line 22 copies the digit into the least significant location. Note the usage of a new constant \textbf{MP\_MASK}. This constant is used to quickly reduce an integer modulo $\beta$. Since $\beta$ is of the form $2^k$ for any suitable $k$ it suffices to perform a binary AND with $MP\_MASK = 2^k - 1$ to perform the reduction. Finally line 23 will set the \textbf{used} member with respect to the digit actually set. This function will always make the integer positive. One important limitation of this function is that it will only set one digit. The size of a digit is not fixed, meaning source that uses this function should take that into account. Meaning that only trivially small constants can be set using this function. \subsection{Setting Large Constants} To overcome the limitations of the mp\_set algorithm the mp\_set\_int algorithm is provided. It accepts a ``long'' data type as input and will always treat it as a 32-bit integer. \begin{figure}[here] \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{mp\_set\_int}. \\ \textbf{Input}. An mp\_int $a$ and a ``long'' integer $b$ \\ \textbf{Output}. Make $a$ equivalent to $b$ \\ \hline \\ 1. Zero $a$ (\textit{mp\_zero}) \\ 2. for $n$ from 0 to 7 do \\ \hspace{3mm}2.1 $a \leftarrow a \cdot 16$ (\textit{mp\_mul2d}) \\ \hspace{3mm}2.2 $u \leftarrow \lfloor b / 2^{4(7 - n)} \rfloor \mbox{ (mod }16\mbox{)}$\\ \hspace{3mm}2.3 $a_0 \leftarrow a_0 + u$ \\ \hspace{3mm}2.4 $a.used \leftarrow a.used + 1$ \\ 3. Clamp excess used digits (\textit{mp\_clamp}) \\ \hline \end{tabular} \end{center} \caption{Algorithm mp\_set\_int} \end{figure} \textbf{Algorithm mp\_set\_int.} The algorithm performs eight iterations of a simple loop where in each iteration four bits from the source are added to the mp\_int. Step 2.1 will multiply the current result by sixteen making room for four more bits in the less significant positions. In step 2.2 the next four bits from the source are extracted and are added to the mp\_int. The \textbf{used} digit count is incremented to reflect the addition. The \textbf{used} digit counter is incremented since if any of the leading digits were zero the mp\_int would have zero digits used and the newly added four bits would be ignored. Excess zero digits are trimmed in steps 2.1 and 3 by using higher level algorithms mp\_mul2d and mp\_clamp. \index{bn\_mp\_set\_int.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_mp\_set\_int.c \vspace{-3mm} \begin{alltt} 016 017 /* set a 32-bit const */ 018 int 019 mp_set_int (mp_int * a, unsigned int b) 020 \{ 021 int x, res; 022 023 mp_zero (a); 024 /* set four bits at a time */ 025 for (x = 0; x < 8; x++) \{ 026 /* shift the number up four bits */ 027 if ((res = mp_mul_2d (a, 4, a)) != MP_OKAY) \{ 028 return res; 029 \} 030 031 /* OR in the top four bits of the source */ 032 a->dp[0] |= (b >> 28) & 15; 033 034 /* shift the source up to the next four bits */ 035 b <<= 4; 036 037 /* ensure that digits are not clamped off */ 038 a->used += 1; 039 \} 040 mp_clamp (a); 041 return MP_OKAY; 042 \} \end{alltt} \end{small} This function sets four bits of the number at a time to handle all practical \textbf{DIGIT\_BIT} sizes. The weird addition on line 38 ensures that the newly added in bits are added to the number of digits. While it may not seem obvious as to why the digit counter does not grow exceedingly large it is because of the shift on line 27 as well as the call to mp\_clamp() on line 40. Both functions will clamp excess leading digits which keeps the number of used digits low. \section{Comparisons} \subsection{Unsigned Comparisions} Comparing a multiple precision integer is performed with the exact same algorithm used to compare two decimal numbers. For example, to compare $1,234$ to $1,264$ the digits are extracted by their positions. That is we compare $1 \cdot 10^3 + 2 \cdot 10^2 + 3 \cdot 10^1 + 4 \cdot 10^0$ to $1 \cdot 10^3 + 2 \cdot 10^2 + 6 \cdot 10^1 + 4 \cdot 10^0$ by comparing single digits at a time starting with the highest magnitude positions. If any leading digit of one integer is greater than a digit in the same position of another integer then obviously it must be greater. The first comparision routine that will be developed is the unsigned magnitude compare which will perform a comparison based on the digits of two mp\_int variables alone. It will ignore the sign of the two inputs. Such a function is useful when an absolute comparison is required or if the signs are known to agree in advance. To facilitate working with the results of the comparison functions three constants are required. \begin{figure}[here] \begin{center} \begin{tabular}{|r|l|} \hline \textbf{Constant} & \textbf{Meaning} \\ \hline \textbf{MP\_GT} & Greater Than \\ \hline \textbf{MP\_EQ} & Equal To \\ \hline \textbf{MP\_LT} & Less Than \\ \hline \end{tabular} \end{center} \caption{Comparison Return Codes} \end{figure} \begin{figure}[here] \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{mp\_cmp\_mag}. \\ \textbf{Input}. Two mp\_ints $a$ and $b$. \\ \textbf{Output}. Unsigned comparison results ($a$ to the left of $b$). \\ \hline \\ 1. If $a.used > b.used$ then return(\textit{MP\_GT}) \\ 2. If $a.used < b.used$ then return(\textit{MP\_LT}) \\ 3. for n from $a.used - 1$ to 0 do \\ \hspace{+3mm}3.1 if $a_n > b_n$ then return(\textit{MP\_GT}) \\ \hspace{+3mm}3.2 if $a_n < b_n$ then return(\textit{MP\_LT}) \\ 4. Return(\textit{MP\_EQ}) \\ \hline \end{tabular} \end{center} \caption{Algorithm mp\_cmp\_mag} \end{figure} \textbf{Algorithm mp\_cmp\_mag.} By saying ``$a$ to the left of $b$'' it is meant that the comparison is with respect to $a$, that is if $a$ is greater than $b$ it will return \textbf{MP\_GT} and similar with respect to when $a = b$ and $a < b$. The first two steps compare the number of digits used in both $a$ and $b$. Obviously if the digit counts differ there would be an imaginary zero digit in the smaller number where the leading digit of the larger number is. If both have the same number of digits than the actual digits themselves must be compared starting at the leading digit. By step three both inputs must have the same number of digits so its safe to start from either $a.used - 1$ or $b.used - 1$ and count down to the zero'th digit. If after all of the digits have been compared, no difference is found, the algorithm returns \textbf{MP\_EQ}. \index{bn\_mp\_cmp\_mag.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_mp\_cmp\_mag.c \vspace{-3mm} \begin{alltt} 016 017 /* compare maginitude of two ints (unsigned) */ 018 int 019 mp_cmp_mag (mp_int * a, mp_int * b) 020 \{ 021 int n; 022 023 /* compare based on # of non-zero digits */ 024 if (a->used > b->used) \{ 025 return MP_GT; 026 \} 027 028 if (a->used < b->used) \{ 029 return MP_LT; 030 \} 031 032 /* compare based on digits */ 033 for (n = a->used - 1; n >= 0; n--) \{ 034 if (a->dp[n] > b->dp[n]) \{ 035 return MP_GT; 036 \} 037 038 if (a->dp[n] < b->dp[n]) \{ 039 return MP_LT; 040 \} 041 \} 042 return MP_EQ; 043 \} \end{alltt} \end{small} The two if statements on lines 24 and 28 compare the number of digits in the two inputs. These two are performed before all of the digits are compared since it is a very cheap test to perform and can potentially save considerable time. The implementation given is also not valid without those two statements. $b.alloc$ may be smaller than $a.used$, meaning that undefined values will be read from $b$ past the end of the array of digits. \subsection{Signed Comparisons} Comparing with sign considerations is also fairly critical in several routines (\textit{division for example}). Based on an unsigned magnitude comparison a trivial signed comparison algorithm can be written. \begin{figure}[here] \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{mp\_cmp}. \\ \textbf{Input}. Two mp\_ints $a$ and $b$ \\ \textbf{Output}. Signed Comparison Results ($a$ to the left of $b$) \\ \hline \\ 1. if $a.sign = MP\_NEG$ and $b.sign = MP\_ZPOS$ then return(\textit{MP\_LT}) \\ 2. if $a.sign = MP\_ZPOS$ and $b.sign = MP\_NEG$ then return(\textit{MP\_GT}) \\ 3. if $a.sign = MP\_NEG$ then \\ \hspace{+3mm}3.1 Return the unsigned comparison of $b$ and $a$ (\textit{mp\_cmp\_mag}) \\ 4 Otherwise \\ \hspace{+3mm}4.1 Return the unsigned comparison of $a$ and $b$ \\ \hline \end{tabular} \end{center} \caption{Algorithm mp\_cmp} \end{figure} \textbf{Algorithm mp\_cmp.} The first two steps compare the signs of the two inputs. If the signs do not agree then it can return right away with the appropriate comparison code. When the signs are equal the digits of the inputs must be compared to determine the correct result. In step three the unsigned comparision flips the order of the arguments since they are both negative. For instance, if $-a > -b$ then $\vert a \vert < \vert b \vert$. Step number four will compare the two when they are both positive. \index{bn\_mp\_cmp.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_mp\_cmp.c \vspace{-3mm} \begin{alltt} 016 017 /* compare two ints (signed)*/ 018 int 019 mp_cmp (mp_int * a, mp_int * b) 020 \{ 021 /* compare based on sign */ 022 if (a->sign == MP_NEG && b->sign == MP_ZPOS) \{ 023 return MP_LT; 024 \} 025 026 if (a->sign == MP_ZPOS && b->sign == MP_NEG) \{ 027 return MP_GT; 028 \} 029 030 /* compare digits */ 031 if (a->sign == MP_NEG) \{ 032 /* if negative compare opposite direction */ 033 return mp_cmp_mag(b, a); 034 \} else \{ 035 return mp_cmp_mag(a, b); 036 \} 037 \} \end{alltt} \end{small} The two if statements on lines 22 and 26 perform the initial sign comparison. If the signs are not the equal then which ever has the positive sign is larger. At line 31, the inputs are compared based on magnitudes. If the signs were both negative then the unsigned comparison is performed in the opposite direction (\textit{line 33}). Otherwise, the signs are assumed to be both positive and a forward direction unsigned comparison is performed. \section*{Exercises} \begin{tabular}{cl} $\left [ 2 \right ]$ & Modify algorithm mp\_set\_int to accept as input a variable length array of bits. \\ & \\ $\left [ 3 \right ]$ & Give the probability that algorithm mp\_cmp\_mag will have to compare $k$ digits \\ & of two random digits (of equal magnitude) before a difference is found. \\ & \\ $\left [ 1 \right ]$ & Suggest a simple method to speed up the implementation of mp\_cmp\_mag based \\ & on the observations made in the previous problem. \\ & \end{tabular} \chapter{Basic Arithmetic} \section{Building Blocks} At this point algorithms for initialization, clearing, zeroing, copying, comparing and setting small constants have been established. The next logical set of algorithms to develop are addition, subtraction and digit shifting algorithms. These algorithms make use of the lower level algorithms and are the cruicial building block for the multiplication algorithms. It is very important that these algorithms are highly optimized. On their own they are simple $O(n)$ algorithms but they can be called from higher level algorithms which easily places them at $O(n^2)$ or even $O(n^3)$ work levels. All nine algorithms within this chapter make use of the logical bit shift operations denoted by $<<$ and $>>$ for left and right logical shifts respectively. A logical shift is analogous to sliding the decimal point of radix-10 representations. For example, the real number $0.9345$ is equivalent to $93.45\%$ which is found by sliding the the decimal two places to the right (\textit{multiplying by $10^2$}). Mathematically a logical shift is equivalent to a division or multiplication by a power of two. For example, $a << k = a \cdot 2^k$ while $a >> k = \lfloor a/2^k \rfloor$. One significant difference between a logical shift and the way decimals are shifted is that digits below the zero'th position are removed from the number. For example, consider $1101_2 >> 1$ using decimal notation this would produce $110.1_2$. However, with a logical shift the result is $110_2$. \section{Addition and Subtraction} In normal fixed precision arithmetic negative numbers are easily represented by subtraction from the modulus. For example, with 32-bit integers $a - b\mbox{ (mod }2^{32}\mbox{)}$ is the same as $a + (2^{32} - b) \mbox{ (mod }2^{32}\mbox{)}$ since $2^{32} \equiv 0 \mbox{ (mod }2^{32}\mbox{)}$. As a result subtraction can be performed with a trivial series of logical operations and an addition. However, in multiple precision arithmetic negative numbers are not represented in the same way. Instead a sign flag is used to keep track of the sign of the integer. As a result signed addition and subtraction are actually implemented as conditional usage of lower level addition or subtraction algorithms with the sign fixed up appropriately. The lower level algorithms will add or subtract integers without regard to the sign flag. That is they will add or subtract the magnitude of the integers respectively. \subsection{Low Level Addition} An unsigned addition of multiple precision integers is performed with the same long-hand algorithm used to add decimal numbers. That is to add the trailing digits first and propagate the resulting carry upwards. Since this is a lower level algorithm the name will have a ``s\_'' prefix. Historically that convention stems from the MPI library where ``s\_'' stood for static functions that were hidden from the developer entirely. \newpage \begin{figure}[!here] \begin{center} \begin{small} \begin{tabular}{l} \hline Algorithm \textbf{s\_mp\_add}. \\ \textbf{Input}. Two mp\_ints $a$ and $b$ \\ \textbf{Output}. The unsigned addition $c = \vert a \vert + \vert b \vert$. \\ \hline \\ 1. if $a.used > b.used$ then \\ \hspace{+3mm}1.1 $min \leftarrow b.used$ \\ \hspace{+3mm}1.2 $max \leftarrow a.used$ \\ \hspace{+3mm}1.3 $x \leftarrow a$ \\ 2. else \\ \hspace{+3mm}2.1 $min \leftarrow a.used$ \\ \hspace{+3mm}2.2 $max \leftarrow b.used$ \\ \hspace{+3mm}2.3 $x \leftarrow b$ \\ 3. If $c.alloc < max + 1$ then grow $c$ to hold at least $max + 1$ digits (\textit{mp\_grow}) \\ 4. If failed to grow $c$ return(\textit{MP\_MEM}) \\ 5. $oldused \leftarrow c.used$ \\ 6. $c.used \leftarrow max + 1$ \\ 7. $u \leftarrow 0$ \\ 8. for $n$ from $0$ to $min - 1$ do \\ \hspace{+3mm}8.1 $c_n \leftarrow a_n + b_n + u$ \\ \hspace{+3mm}8.2 $u \leftarrow c_n >> lg(\beta)$ \\ \hspace{+3mm}8.3 $c_n \leftarrow c_n \mbox{ (mod }\beta\mbox{)}$ \\ 9. if $min \ne max$ then do \\ \hspace{+3mm}9.1 for $n$ from $min$ to $max - 1$ do \\ \hspace{+6mm}9.1.1 $c_n \leftarrow x_n + u$ \\ \hspace{+6mm}9.1.2 $u \leftarrow c_n >> lg(\beta)$ \\ \hspace{+6mm}9.1.3 $c_n \leftarrow c_n \mbox{ (mod }\beta\mbox{)}$ \\ 10. $c_{max} \leftarrow u$ \\ 11. if $olduse > max$ then \\ \hspace{+3mm}11.1 for $n$ from $max + 1$ to $olduse - 1$ do \\ \hspace{+6mm}11.1.1 $c_n \leftarrow 0$ \\ 12. Clamp excess digits in $c$. (\textit{mp\_clamp}) \\ 13. Return(\textit{MP\_OKAY}) \\ \hline \end{tabular} \end{small} \end{center} \caption{Algorithm s\_mp\_add} \end{figure} \textbf{Algorithm s\_mp\_add.} This algorithm is loosely based on algorithm 14.7 of HAC \cite[pp. 594]{HAC} but has been extended to allow the inputs to have different magnitudes. Coincidentally the description of algorithm A in Knuth \cite[pp. 266]{TAOCPV2} shares the same deficiency as the algorithm from \cite{HAC}. Even the MIX pseudo machine code presented by Knuth \cite[pp. 266-267]{TAOCPV2} is incapable of handling inputs which are of different magnitudes. Steps 1 and 2 will sort the two inputs based on their \textbf{used} digit count. This allows the inputs to have varying magnitudes which not only makes it more efficient than the trivial algorithm presented in the references but more flexible. The variable $min$ is given the lowest digit count while $max$ is given the highest digit count. If both inputs have the same \textbf{used} digit count both $min$ and $max$ are set to the same value. The variable $x$ is an \textit{alias} for the largest input and not meant to be a copy of it. After the inputs are sorted, steps 3 and 4 will ensure that the destination $c$ can accommodate the result. The old \textbf{used} count from $c$ is copied to $oldused$ so that excess digits can be cleared later, and the new \textbf{used} count is set to $max+1$, so that a carry from the most significant word can be handled. At step 7 the carry variable $u$ is set to zero and the first part of the addition loop can begin. The first step of the loop (\textit{8.1}) adds digits from the two inputs together along with the carry variable $u$. The following step extracts the carry bit by shifting the result of the preceding step right by $lg(\beta)$ positions. The shift to extract the carry is similar to how carry extraction works with decimal addition. Consider adding $77$ to $65$, the first addition of the first column is $7 + 5$ which produces the result $12$. The trailing digit of the result is $2 \equiv 12 \mbox{ (mod }10\mbox{)}$ and the carry is found by dividing (\textit{and ignoring the remainder}) $12$ by the radix or in this case $10$. The division and multiplication of $10$ is simply a logical right or left shift, respectively, of the digits. In otherwords the carry can be extracted by shifting one digit to the right. Note that $lg()$ is simply the base two logarithm such that $lg(2^k) = k$. This implies that $lg(\beta)$ is the number of bits in a radix-$\beta$ digit. Therefore, a logical shift right of the summand by $lg(\beta)$ will extract the carry. The final step of the loop reduces the digit modulo the radix $\beta$ to ensure it is in range. After step 8 the smallest input (\textit{or both if they are the same magnitude}) has been exhausted. Step 9 decides whether the inputs were of equal magnitude. If not than another loop similar to that in step 8, must be executed. The loop at step number 9.1 differs from the previous loop since it only adds the mp\_int $x$ along with the carry. Step 10 finishes the addition phase by copying the final carry to the highest location in the result $c_{max}$. Step 11 ensures that leading digits that were originally present in $c$ are cleared. Finally excess leading digits are clamped and the algorithm returns success. \index{bn\_s\_mp\_add.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_s\_mp\_add.c \vspace{-3mm} \begin{alltt} 016 017 /* low level addition, based on HAC pp.594, Algorithm 14.7 */ 018 int 019 s_mp_add (mp_int * a, mp_int * b, mp_int * c) 020 \{ 021 mp_int *x; 022 int olduse, res, min, max; 023 024 /* find sizes, we let |a| <= |b| which means we have to sort 025 * them. "x" will point to the input with the most digits 026 */ 027 if (a->used > b->used) \{ 028 min = b->used; 029 max = a->used; 030 x = a; 031 \} else \{ 032 min = a->used; 033 max = b->used; 034 x = b; 035 \} 036 037 /* init result */ 038 if (c->alloc < max + 1) \{ 039 if ((res = mp_grow (c, max + 1)) != MP_OKAY) \{ 040 return res; 041 \} 042 \} 043 044 /* get old used digit count and set new one */ 045 olduse = c->used; 046 c->used = max + 1; 047 048 \{ 049 register mp_digit u, *tmpa, *tmpb, *tmpc; 050 register int i; 051 052 /* alias for digit pointers */ 053 054 /* first input */ 055 tmpa = a->dp; 056 057 /* second input */ 058 tmpb = b->dp; 059 060 /* destination */ 061 tmpc = c->dp; 062 063 /* zero the carry */ 064 u = 0; 065 for (i = 0; i < min; i++) \{ 066 /* Compute the sum at one digit, T[i] = A[i] + B[i] + U */ 067 *tmpc = *tmpa++ + *tmpb++ + u; 068 069 /* U = carry bit of T[i] */ 070 u = *tmpc >> ((mp_digit)DIGIT_BIT); 071 072 /* take away carry bit from T[i] */ 073 *tmpc++ &= MP_MASK; 074 \} 075 076 /* now copy higher words if any, that is in A+B 077 * if A or B has more digits add those in 078 */ 079 if (min != max) \{ 080 for (; i < max; i++) \{ 081 /* T[i] = X[i] + U */ 082 *tmpc = x->dp[i] + u; 083 084 /* U = carry bit of T[i] */ 085 u = *tmpc >> ((mp_digit)DIGIT_BIT); 086 087 /* take away carry bit from T[i] */ 088 *tmpc++ &= MP_MASK; 089 \} 090 \} 091 092 /* add carry */ 093 *tmpc++ = u; 094 095 /* clear digits above oldused */ 096 for (i = c->used; i < olduse; i++) \{ 097 *tmpc++ = 0; 098 \} 099 \} 100 101 mp_clamp (c); 102 return MP_OKAY; 103 \} \end{alltt} \end{small} Lines 27 to 35 perform the initial sorting of the inputs and determine the $min$ and $max$ variables. Note that $x$ is a pointer to a mp\_int assigned to the largest input, in effect it is a local alias. Lines 37 to 42 ensure that the destination is grown to accomodate the result of the addition. Similar to the implementation of mp\_copy this function uses the braced code and local aliases coding style. The three aliases that are on lines 55, 58 and 61 represent the two inputs and destination variables respectively. These aliases are used to ensure the compiler does not have to dereference $a$, $b$ or $c$ (respectively) to access the digits of the respective mp\_int. The initial carry $u$ is cleared on line 64, note that $u$ is of type mp\_digit which ensures type compatibility within the implementation. The initial addition loop begins on line 65 and ends on line 74. Similarly the conditional addition loop begins on line 80 and ends on line 90. The addition is finished with the final carry being stored in $tmpc$ on line 93. Note the ``++'' operator on the same line. After line 93 $tmpc$ will point to the $c.used$'th digit of the mp\_int $c$. This is useful for the next loop on lines 96 to 99 which set any old upper digits to zero. \subsection{Low Level Subtraction} The low level unsigned subtraction algorithm is very similar to the low level unsigned addition algorithm. The principle difference is that the unsigned subtraction algorithm requires the result to be positive. That is when computing $a - b$ the condition $\vert a \vert \ge \vert b\vert$ must be met for this algorithm to function properly. Keep in mind this low level algorithm is not meant to be used in higher level algorithms directly. This algorithm as will be shown can be used to create functional signed addition and subtraction algorithms. For this algorithm a new variable is required to make the description simpler. Recall from section 1.3.1 that a mp\_digit must be able to represent the range $0 \le x < 2\beta$ for the algorithms to work correctly. However, it is allowable that a mp\_digit represent a larger range of values. For this algorithm we will assume that the variable $\gamma$ represents the number of bits available in a mp\_digit (\textit{this implies $2^{\gamma} > \beta$}). For example, the default for LibTomMath is to use a ``unsigned long'' for the mp\_digit ``type'' while $\beta = 2^{28}$. In ISO C an ``unsigned long'' data type must be able to represent $0 \le x < 2^{32}$ meaning that in this case $\gamma = 32$. \newpage\begin{figure}[!here] \begin{center} \begin{small} \begin{tabular}{l} \hline Algorithm \textbf{s\_mp\_sub}. \\ \textbf{Input}. Two mp\_ints $a$ and $b$ ($\vert a \vert \ge \vert b \vert$) \\ \textbf{Output}. The unsigned subtraction $c = \vert a \vert - \vert b \vert$. \\ \hline \\ 1. $min \leftarrow b.used$ \\ 2. $max \leftarrow a.used$ \\ 3. If $c.alloc < max$ then grow $c$ to hold at least $max$ digits. (\textit{mp\_grow}) \\ 4. If the reallocation failed return(\textit{MP\_MEM}). \\ 5. $oldused \leftarrow c.used$ \\ 6. $c.used \leftarrow max$ \\ 7. $u \leftarrow 0$ \\ 8. for $n$ from $0$ to $min - 1$ do \\ \hspace{3mm}8.1 $c_n \leftarrow a_n - b_n - u$ \\ \hspace{3mm}8.2 $u \leftarrow c_n >> (\gamma - 1)$ \\ \hspace{3mm}8.3 $c_n \leftarrow c_n \mbox{ (mod }\beta\mbox{)}$ \\ 9. if $min < max$ then do \\ \hspace{3mm}9.1 for $n$ from $min$ to $max - 1$ do \\ \hspace{6mm}9.1.1 $c_n \leftarrow a_n - u$ \\ \hspace{6mm}9.1.2 $u \leftarrow c_n >> (\gamma - 1)$ \\ \hspace{6mm}9.1.3 $c_n \leftarrow c_n \mbox{ (mod }\beta\mbox{)}$ \\ 10. if $oldused > max$ then do \\ \hspace{3mm}10.1 for $n$ from $max$ to $oldused - 1$ do \\ \hspace{6mm}10.1.1 $c_n \leftarrow 0$ \\ 11. Clamp excess digits of $c$. (\textit{mp\_clamp}). \\ 12. Return(\textit{MP\_OKAY}). \\ \hline \end{tabular} \end{small} \end{center} \caption{Algorithm s\_mp\_sub} \end{figure} \textbf{Algorithm s\_mp\_sub.} This algorithm performs the unsigned subtraction of two mp\_int variables under the restriction that the result must be positive. That is when passing variables $a$ and $b$ the condition that $\vert a \vert \ge \vert b \vert$ must be met for the algorithm to function correctly. This algorithm is loosely based on algorithm 14.9 \cite[pp. 595]{HAC} and is similar to algorithm S in \cite[pp. 267]{TAOCPV2} as well. As was the case of the algorithm s\_mp\_add both other references lack discussion concerning various practical details such as when the inputs differ in magnitude. The initial sorting of the inputs is trivial in this algorithm since $a$ is guaranteed to have at least the same magnitude of $b$. Steps 1 and 2 set the $min$ and $max$ variables. Unlike the addition routine there is guaranteed to be no carry which means that the final result can be at most $max$ digits in length as opposed to $max + 1$. Similar to the addition algorithm the \textbf{used} count of $c$ is copied locally and set to the maximal count for the operation. The subtraction loop that begins on step 8 is essentially the same as the addition loop of algorithm s\_mp\_add except single precision subtraction is used instead. Note the use of the $\gamma$ variable to extract the carry (\textit{also known as the borrow}) within the subtraction loops. Under the assumption that two's complement single precision arithmetic is used this will successfully extract the desired carry. For example, consider subtracting $0101_2$ from $0100_2$ where $\gamma = 4$ and $\beta = 2$. The least significant bit will force a carry upwards to the third bit which will be set to zero after the borrow. After the very first bit has been subtracted $4 - 1 \equiv 0011_2$ will remain, When the third bit of $0101_2$ is subtracted from the result it will cause another carry. In this case though the carry will be forced to propagate all the way to the most significant bit. Recall that $\beta < 2^{\gamma}$. This means that if a carry does occur just before the $lg(\beta)$'th bit it will propagate all the way to the most significant bit. Thus, the high order bits of the mp\_digit that are not part of the actual digit will either be all zero, or all one. All that is needed is a single zero or one bit for the carry. Therefore a single logical shift right by $\gamma - 1$ positions is sufficient to extract the carry. This method of carry extraction may seem awkward but the reason for it becomes apparent when the implementation is discussed. If $b$ has a smaller magnitude than $a$ then step 9 will force the carry and copy operation to propagate through the larger input $a$ into $c$. Step 10 will ensure that any leading digits of $c$ above the $max$'th position are zeroed. \index{bn\_s\_mp\_sub.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_s\_mp\_sub.c \vspace{-3mm} \begin{alltt} 016 017 /* low level subtraction (assumes |a| > |b|), HAC pp.595 Algorithm 14.9 */ 018 int 019 s_mp_sub (mp_int * a, mp_int * b, mp_int * c) 020 \{ 021 int olduse, res, min, max; 022 023 /* find sizes */ 024 min = b->used; 025 max = a->used; 026 027 /* init result */ 028 if (c->alloc < max) \{ 029 if ((res = mp_grow (c, max)) != MP_OKAY) \{ 030 return res; 031 \} 032 \} 033 olduse = c->used; 034 c->used = max; 035 036 \{ 037 register mp_digit u, *tmpa, *tmpb, *tmpc; 038 register int i; 039 040 /* alias for digit pointers */ 041 tmpa = a->dp; 042 tmpb = b->dp; 043 tmpc = c->dp; 044 045 /* set carry to zero */ 046 u = 0; 047 for (i = 0; i < min; i++) \{ 048 /* T[i] = A[i] - B[i] - U */ 049 *tmpc = *tmpa++ - *tmpb++ - u; 050 051 /* U = carry bit of T[i] 052 * Note this saves performing an AND operation since 053 * if a carry does occur it will propagate all the way to the 054 * MSB. As a result a single shift is enough to get the carry 055 */ 056 u = *tmpc >> ((mp_digit)(CHAR_BIT * sizeof (mp_digit) - 1)); 057 058 /* Clear carry from T[i] */ 059 *tmpc++ &= MP_MASK; 060 \} 061 062 /* now copy higher words if any, e.g. if A has more digits than B */ 063 for (; i < max; i++) \{ 064 /* T[i] = A[i] - U */ 065 *tmpc = *tmpa++ - u; 066 067 /* U = carry bit of T[i] */ 068 u = *tmpc >> ((mp_digit)(CHAR_BIT * sizeof (mp_digit) - 1)); 069 070 /* Clear carry from T[i] */ 071 *tmpc++ &= MP_MASK; 072 \} 073 074 /* clear digits above used (since we may not have grown result above) */ 075 for (i = c->used; i < olduse; i++) \{ 076 *tmpc++ = 0; 077 \} 078 \} 079 080 mp_clamp (c); 081 return MP_OKAY; 082 \} 083 \end{alltt} \end{small} Line 24 and 25 perform the initial hardcoded sorting of the inputs. In reality the $min$ and $max$ variables are only aliases and are only used to make the source code easier to read. Again the pointer alias optimization is used within this algorithm. Lines 41, 42 and 43 initialize the aliases for $a$, $b$ and $c$ respectively. The first subtraction loop occurs on lines 46 through 60. The theory behind the subtraction loop is exactly the same as that for the addition loop. As remarked earlier there is an implementation reason for using the ``awkward'' method of extracting the carry (\textit{see line 56}). The traditional method for extracting the carry would be to shift by $lg(\beta)$ positions and logically AND the least significant bit. The AND operation is required because all of the bits above the $\lg(\beta)$'th bit will be set to one after a carry occurs from subtraction. This carry extraction requires two relatively cheap operations to extract the carry. The other method is to simply shift the most significant bit to the least significant bit thus extracting the carry with a single cheap operation. This optimization only works on twos compliment machines which is a safe assumption to make. If $a$ has a larger magnitude than $b$ an additional loop (\textit{see lines 63 through 72}) is required to propagate the carry through $a$ and copy the result to $c$. \subsection{High Level Addition} Now that both lower level addition and subtraction algorithms have been established an effective high level signed addition algorithm can be established. This high level addition algorithm will be what other algorithms and developers will use to perform addition of mp\_int data types. Recall from section 5.2 that an mp\_int represents an integer with an unsigned mantissa (\textit{the array of digits}) and a \textbf{sign} flag. A high level addition is actually performed as a series of eight separate cases which can be optimized down to three unique cases. \begin{figure}[!here] \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{mp\_add}. \\ \textbf{Input}. Two mp\_ints $a$ and $b$ \\ \textbf{Output}. The signed addition $c = a + b$. \\ \hline \\ 1. if $a.sign = b.sign$ then do \\ \hspace{3mm}1.1 $c.sign \leftarrow a.sign$ \\ \hspace{3mm}1.2 $c \leftarrow \vert a \vert + \vert b \vert$ (\textit{s\_mp\_add})\\ 2. else do \\ \hspace{3mm}2.1 if $\vert a \vert < \vert b \vert$ then do (\textit{mp\_cmp\_mag}) \\ \hspace{6mm}2.1.1 $c.sign \leftarrow b.sign$ \\ \hspace{6mm}2.1.2 $c \leftarrow \vert b \vert - \vert a \vert$ (\textit{s\_mp\_sub}) \\ \hspace{3mm}2.2 else do \\ \hspace{6mm}2.2.1 $c.sign \leftarrow a.sign$ \\ \hspace{6mm}2.2.2 $c \leftarrow \vert a \vert - \vert b \vert$ \\ 3. If any of the lower level operations failed return(\textit{MP\_MEM}) \\ 4. Return(\textit{MP\_OKAY}). \\ \hline \end{tabular} \end{center} \caption{Algorithm mp\_add} \end{figure} \textbf{Algorithm mp\_add.} This algorithm performs the signed addition of two mp\_int variables. There is no reference algorithm to draw upon from either \cite{TAOCPV2} or \cite{HAC} since they both only provide unsigned operations. The algorithm is fairly straightforward but restricted since subtraction can only produce positive results. \begin{figure}[here] \begin{small} \begin{center} \begin{tabular}{|c|c|c|c|c|} \hline \textbf{Sign of $a$} & \textbf{Sign of $b$} & \textbf{$\vert a \vert > \vert b \vert $} & \textbf{Unsigned Operation} & \textbf{Result Sign Flag} \\ \hline $+$ & $+$ & Yes & $c = a + b$ & $a.sign$ \\ \hline $+$ & $+$ & No & $c = a + b$ & $a.sign$ \\ \hline $-$ & $-$ & Yes & $c = a + b$ & $a.sign$ \\ \hline $-$ & $-$ & No & $c = a + b$ & $a.sign$ \\ \hline &&&&\\ \hline $+$ & $-$ & No & $c = b - a$ & $b.sign$ \\ \hline $-$ & $+$ & No & $c = b - a$ & $b.sign$ \\ \hline &&&&\\ \hline $+$ & $-$ & Yes & $c = a - b$ & $a.sign$ \\ \hline $-$ & $+$ & Yes & $c = a - b$ & $a.sign$ \\ \hline \end{tabular} \end{center} \end{small} \caption{Addition Guide Chart} \label{fig:AddChart} \end{figure} Figure~\ref{fig:AddChart} lists all of the eight possible input combinations and is sorted to show that only three specific cases need to be handled. The return code of the unsigned operations at step 1.2, 2.1.2 and 2.2.2 are forwarded to step 3 to check for errors. This simplifies the description of the algorithm considerably and best follows how the implementation actually was achieved. Also note how the \textbf{sign} is set before the unsigned addition or subtraction is performed. Recall from the descriptions of algorithms s\_mp\_add and s\_mp\_sub that the mp\_clamp function is used at the end to trim excess digits. The mp\_clamp algorithm will set the \textbf{sign} to \textbf{MP\_ZPOS} when the \textbf{used} digit count reaches zero. For example, consider performing $-a + a$ with algorithm mp\_add. By the description of the algorithm the sign is set to \textbf{MP\_NEG} which would produce a result of $-0$. However, since the sign is set first then the unsigned addition is performed the subsequent usage of algorithm mp\_clamp within algorithm s\_mp\_add will force $-0$ to become $0$. \index{bn\_mp\_add.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_mp\_add.c \vspace{-3mm} \begin{alltt} 016 017 /* high level addition (handles signs) */ 018 int 019 mp_add (mp_int * a, mp_int * b, mp_int * c) 020 \{ 021 int sa, sb, res; 022 023 /* get sign of both inputs */ 024 sa = a->sign; 025 sb = b->sign; 026 027 /* handle two cases, not four */ 028 if (sa == sb) \{ 029 /* both positive or both negative */ 030 /* add their magnitudes, copy the sign */ 031 c->sign = sa; 032 res = s_mp_add (a, b, c); 033 \} else \{ 034 /* one positive, the other negative */ 035 /* subtract the one with the greater magnitude from */ 036 /* the one of the lesser magnitude. The result gets */ 037 /* the sign of the one with the greater magnitude. */ 038 if (mp_cmp_mag (a, b) == MP_LT) \{ 039 c->sign = sb; 040 res = s_mp_sub (b, a, c); 041 \} else \{ 042 c->sign = sa; 043 res = s_mp_sub (a, b, c); 044 \} 045 \} 046 return res; 047 \} 048 \end{alltt} \end{small} The source code follows the algorithm fairly closely. The most notable new source code addition is the usage of the $res$ integer variable which is used to pass result of the unsigned operations forward. Unlike in the algorithm, the variable $res$ is merely returned as is without explicitly checking it and returning the constant \textbf{MP\_OKAY}. The observation is this algorithm will succeed or fail only if the lower level functions do so. Returning their return code is sufficient. \subsection{High Level Subtraction} The high level signed subtraction algorithm is essentially the same as the high level signed addition algorithm. \newpage\begin{figure}[!here] \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{mp\_sub}. \\ \textbf{Input}. Two mp\_ints $a$ and $b$ \\ \textbf{Output}. The signed subtraction $c = a - b$. \\ \hline \\ 1. if $a.sign \ne b.sign$ then do \\ \hspace{3mm}1.1 $c.sign \leftarrow a.sign$ \\ \hspace{3mm}1.2 $c \leftarrow \vert a \vert + \vert b \vert$ (\textit{s\_mp\_add}) \\ 2. else do \\ \hspace{3mm}2.1 if $\vert a \vert \ge \vert b \vert$ then do (\textit{mp\_cmp\_mag}) \\ \hspace{6mm}2.1.1 $c.sign \leftarrow a.sign$ \\ \hspace{6mm}2.1.2 $c \leftarrow \vert a \vert - \vert b \vert$ (\textit{s\_mp\_sub}) \\ \hspace{3mm}2.2 else do \\ \hspace{6mm}2.2.1 $c.sign \leftarrow \left \lbrace \begin{array}{ll} MP\_ZPOS & \mbox{if }a.sign = MP\_NEG \\ MP\_NEG & \mbox{otherwise} \\ \end{array} \right .$ \\ \hspace{6mm}2.2.2 $c \leftarrow \vert b \vert - \vert a \vert$ \\ 3. If any of the lower level operations failed return(\textit{MP\_MEM}). \\ 4. Return(\textit{MP\_OKAY}). \\ \hline \end{tabular} \end{center} \caption{Algorithm mp\_sub} \end{figure} \textbf{Algorithm mp\_sub.} This algorithm performs the signed subtraction of two inputs. Similar to algorithm mp\_add there is no reference in either \cite{TAOCPV2} or \cite{HAC}. Also this algorithm is restricted by algorithm s\_mp\_sub. The following chart lists the eight possible inputs and the operations required. \begin{figure}[!here] \begin{small} \begin{center} \begin{tabular}{|c|c|c|c|c|} \hline \textbf{Sign of $a$} & \textbf{Sign of $b$} & \textbf{$\vert a \vert \ge \vert b \vert $} & \textbf{Unsigned Operation} & \textbf{Result Sign Flag} \\ \hline $+$ & $-$ & Yes & $c = a + b$ & $a.sign$ \\ \hline $+$ & $-$ & No & $c = a + b$ & $a.sign$ \\ \hline $-$ & $+$ & Yes & $c = a + b$ & $a.sign$ \\ \hline $-$ & $+$ & No & $c = a + b$ & $a.sign$ \\ \hline &&&& \\ \hline $+$ & $+$ & Yes & $c = a - b$ & $a.sign$ \\ \hline $-$ & $-$ & Yes & $c = a - b$ & $a.sign$ \\ \hline &&&& \\ \hline $+$ & $+$ & No & $c = b - a$ & $\mbox{opposite of }a.sign$ \\ \hline $-$ & $-$ & No & $c = b - a$ & $\mbox{opposite of }a.sign$ \\ \hline \end{tabular} \end{center} \end{small} \caption{Subtraction Guide Chart} \end{figure} Similar to the case of algorithm mp\_add the \textbf{sign} is set first before the unsigned addition or subtraction. That is to prevent the algorithm from producing $-a - -a = -0$ as a result. \index{bn\_mp\_sub.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_mp\_sub.c \vspace{-3mm} \begin{alltt} 016 017 /* high level subtraction (handles signs) */ 018 int 019 mp_sub (mp_int * a, mp_int * b, mp_int * c) 020 \{ 021 int sa, sb, res; 022 023 sa = a->sign; 024 sb = b->sign; 025 026 if (sa != sb) \{ 027 /* subtract a negative from a positive, OR */ 028 /* subtract a positive from a negative. */ 029 /* In either case, ADD their magnitudes, */ 030 /* and use the sign of the first number. */ 031 c->sign = sa; 032 res = s_mp_add (a, b, c); 033 \} else \{ 034 /* subtract a positive from a positive, OR */ 035 /* subtract a negative from a negative. */ 036 /* First, take the difference between their */ 037 /* magnitudes, then... */ 038 if (mp_cmp_mag (a, b) != MP_LT) \{ 039 /* Copy the sign from the first */ 040 c->sign = sa; 041 /* The first has a larger or equal magnitude */ 042 res = s_mp_sub (a, b, c); 043 \} else \{ 044 /* The result has the *opposite* sign from */ 045 /* the first number. */ 046 c->sign = (sa == MP_ZPOS) ? MP_NEG : MP_ZPOS; 047 /* The second has a larger magnitude */ 048 res = s_mp_sub (b, a, c); 049 \} 050 \} 051 return res; 052 \} 053 \end{alltt} \end{small} Much like the implementation of algorithm mp\_add the variable $res$ is used to catch the return code of the unsigned addition or subtraction operations and forward it to the end of the function. On line 38 the ``not equal to'' \textbf{MP\_LT} expression is used to emulate a ``greater than or equal to'' comparison. \section{Bit and Digit Shifting} It is quite common to think of a multiple precision integer as a polynomial in $x$, that is $y = f(\beta)$ where $f(x) = \sum_{i=0}^{n-1} a_i x^i$. This notation arises within discussion of Montgomery and Diminished Radix Reduction as well as Karatsuba multiplication and squaring. In order to facilitate operations on polynomials in $x$ as above a series of simple ``digit'' algorithms have to be established. That is to shift the digits left or right as well to shift individual bits of the digits left and right. It is important to note that not all ``shift'' operations are on radix-$\beta$ digits. \subsection{Multiplication by Two} In a binary system where the radix is a power of two multiplication by two not only arises often in other algorithms it is a fairly efficient operation to perform. A single precision logical shift left is sufficient to multiply a single digit by two. \newpage\begin{figure}[!here] \begin{small} \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{mp\_mul\_2}. \\ \textbf{Input}. One mp\_int $a$ \\ \textbf{Output}. $b = 2a$. \\ \hline \\ 1. If $b.alloc < a.used + 1$ then grow $b$ to hold $a.used + 1$ digits. (\textit{mp\_grow}) \\ 2. If the reallocation failed return(\textit{MP\_MEM}). \\ 3. $oldused \leftarrow b.used$ \\ 4. $b.used \leftarrow a.used$ \\ 5. $r \leftarrow 0$ \\ 6. for $n$ from 0 to $a.used - 1$ do \\ \hspace{3mm}6.1 $rr \leftarrow a_n >> (lg(\beta) - 1)$ \\ \hspace{3mm}6.2 $b_n \leftarrow (a_n << 1) + r \mbox{ (mod }\beta\mbox{)}$ \\ \hspace{3mm}6.3 $r \leftarrow rr$ \\ 7. If $r \ne 0$ then do \\ \hspace{3mm}7.1 $b_{n + 1} \leftarrow r$ \\ \hspace{3mm}7.2 $b.used \leftarrow b.used + 1$ \\ 8. If $b.used < oldused - 1$ then do \\ \hspace{3mm}8.1 for $n$ from $b.used$ to $oldused - 1$ do \\ \hspace{6mm}8.1.1 $b_n \leftarrow 0$ \\ 9. $b.sign \leftarrow a.sign$ \\ 10. Return(\textit{MP\_OKAY}).\\ \hline \end{tabular} \end{center} \end{small} \caption{Algorithm mp\_mul\_2} \end{figure} \textbf{Algorithm mp\_mul\_2.} This algorithm will quickly multiply a mp\_int by two provided $\beta$ is a power of two. Neither \cite{TAOCPV2} nor \cite{HAC} describe such an algorithm despite the fact it arises often in other algorithms. The algorithm is setup much like the lower level algorithm s\_mp\_add since it is for all intents and purposes equivalent to the operation $b = \vert a \vert + \vert a \vert$. Step 1 and 2 grow the input as required to accomodate the maximum number of \textbf{used} digits in the result. The initial \textbf{used} count is set to $a.used$ at step 4. Only if there is a final carry will the \textbf{used} count require adjustment. Step 6 is an optimization implementation of the addition loop for this specific case. That is since the two values being added together are the same there is no need to perform two reads from the digits of $a$. Step 6.1 performs a single precision shift on the current digit $a_n$ to obtain what will be the carry for the next iteration. Step 6.2 calculates the $n$'th digit of the result as single precision shift of $a_n$ plus the previous carry. Recall from section 5.1 that $a_n << 1$ is equivalent to $a_n \cdot 2$. An iteration of the addition loop is finished with forwarding the carry to the next iteration. Step 7 takes care of any final carry by setting the $a.used$'th digit of the result to the carry and augmenting the \textbf{used} count of $b$. Step 8 clears any leading digits of $b$ in case it originally had a larger magnitude than $a$. \index{bn\_mp\_mul\_2.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_mp\_mul\_2.c \vspace{-3mm} \begin{alltt} 016 017 /* b = a*2 */ 018 int 019 mp_mul_2 (mp_int * a, mp_int * b) 020 \{ 021 int x, res, oldused; 022 023 /* grow to accomodate result */ 024 if (b->alloc < a->used + 1) \{ 025 if ((res = mp_grow (b, a->used + 1)) != MP_OKAY) \{ 026 return res; 027 \} 028 \} 029 030 oldused = b->used; 031 b->used = a->used; 032 033 \{ 034 register mp_digit r, rr, *tmpa, *tmpb; 035 036 /* alias for source */ 037 tmpa = a->dp; 038 039 /* alias for dest */ 040 tmpb = b->dp; 041 042 /* carry */ 043 r = 0; 044 for (x = 0; x < a->used; x++) \{ 045 046 /* get what will be the *next* carry bit from the 047 * MSB of the current digit 048 */ 049 rr = *tmpa >> ((mp_digit)(DIGIT_BIT - 1)); 050 051 /* now shift up this digit, add in the carry [from the previous] */ 052 *tmpb++ = ((*tmpa++ << ((mp_digit)1)) | r) & MP_MASK; 053 054 /* copy the carry that would be from the source 055 * digit into the next iteration 056 */ 057 r = rr; 058 \} 059 060 /* new leading digit? */ 061 if (r != 0) \{ 062 /* add a MSB which is always 1 at this point */ 063 *tmpb = 1; 064 ++b->used; 065 \} 066 067 /* now zero any excess digits on the destination 068 * that we didn't write to 069 */ 070 tmpb = b->dp + b->used; 071 for (x = b->used; x < oldused; x++) \{ 072 *tmpb++ = 0; 073 \} 074 \} 075 b->sign = a->sign; 076 return MP_OKAY; 077 \} \end{alltt} \end{small} This implementation is essentially an optimized implementation of s\_mp\_add for the case of doubling an input. The only noteworthy difference is the use of the logical shift operator on line 52 to perform a single precision doubling. \subsection{Division by Two} A division by two can just as easily be accomplished with a logical shift right as multiplication by two can be with a logical shift left. \newpage\begin{figure}[!here] \begin{small} \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{mp\_div\_2}. \\ \textbf{Input}. One mp\_int $a$ \\ \textbf{Output}. $b = a/2$. \\ \hline \\ 1. If $b.alloc < a.used$ then grow $b$ to hold $a.used$ digits. (\textit{mp\_grow}) \\ 2. If the reallocation failed return(\textit{MP\_MEM}). \\ 3. $oldused \leftarrow b.used$ \\ 4. $b.used \leftarrow a.used$ \\ 5. $r \leftarrow 0$ \\ 6. for $n$ from $b.used - 1$ to $0$ do \\ \hspace{3mm}6.1 $rr \leftarrow a_n \mbox{ (mod }2\mbox{)}$\\ \hspace{3mm}6.2 $b_n \leftarrow (a_n >> 1) + (r << (lg(\beta) - 1)) \mbox{ (mod }\beta\mbox{)}$ \\ \hspace{3mm}6.3 $r \leftarrow rr$ \\ 7. If $b.used < oldused - 1$ then do \\ \hspace{3mm}7.1 for $n$ from $b.used$ to $oldused - 1$ do \\ \hspace{6mm}7.1.1 $b_n \leftarrow 0$ \\ 8. $b.sign \leftarrow a.sign$ \\ 9. Clamp excess digits of $b$. (\textit{mp\_clamp}) \\ 10. Return(\textit{MP\_OKAY}).\\ \hline \end{tabular} \end{center} \end{small} \caption{Algorithm mp\_div\_2} \end{figure} \textbf{Algorithm mp\_div\_2.} This algorithm will divide an mp\_int by two using logical shifts to the right. Like mp\_mul\_2 it uses a modified low level addition core as the basis of the algorithm. Unlike mp\_mul\_2 the shift operations work from the leading digit to the trailing digit. The algorithm could be written to work from the trailing digit to the leading digit however, it would have to stop one short of $a.used - 1$ digits to prevent reading past the end of the array of digits. Essentially the loop at step 6 is similar to that of mp\_mul\_2 except the logical shifts go in the opposite direction and the carry is at the least significant bit not the most significant bit. \index{bn\_mp\_div\_2.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_mp\_div\_2.c \vspace{-3mm} \begin{alltt} 016 017 /* b = a/2 */ 018 int 019 mp_div_2 (mp_int * a, mp_int * b) 020 \{ 021 int x, res, oldused; 022 023 /* copy */ 024 if (b->alloc < a->used) \{ 025 if ((res = mp_grow (b, a->used)) != MP_OKAY) \{ 026 return res; 027 \} 028 \} 029 030 oldused = b->used; 031 b->used = a->used; 032 \{ 033 register mp_digit r, rr, *tmpa, *tmpb; 034 035 /* source alias */ 036 tmpa = a->dp + b->used - 1; 037 038 /* dest alias */ 039 tmpb = b->dp + b->used - 1; 040 041 /* carry */ 042 r = 0; 043 for (x = b->used - 1; x >= 0; x--) \{ 044 /* get the carry for the next iteration */ 045 rr = *tmpa & 1; 046 047 /* shift the current digit, add in carry and store */ 048 *tmpb-- = (*tmpa-- >> 1) | (r << (DIGIT_BIT - 1)); 049 050 /* forward carry to next iteration */ 051 r = rr; 052 \} 053 054 /* zero excess digits */ 055 tmpb = b->dp + b->used; 056 for (x = b->used; x < oldused; x++) \{ 057 *tmpb++ = 0; 058 \} 059 \} 060 b->sign = a->sign; 061 mp_clamp (b); 062 return MP_OKAY; 063 \} \end{alltt} \end{small} \section{Polynomial Basis Operations} Recall from section 5.3 that any integer can be represented as a polynomial in $x$ as $y = f(\beta)$. Such a representation is also known as the polynomial basis \cite[pp. 48]{ROSE}. Given such a notation a multiplication or division by $x$ amounts to shifting whole digits a single place. The need for such operations arises in several other higher level algorithms such as Barrett and Montgomery reduction, integer division and Karatsuba multiplication. Converting from an array of digits to polynomial basis is very simple. Consider the integer $y \equiv (a_2, a_1, a_0)_{\beta}$ and recall that $y = \sum_{i=0}^{2} a_i \beta^i$. Simply replace $\beta$ with $x$ and the expression is in polynomial basis. For example, $f(x) = 8x + 9$ is the polynomial basis representation for $89$ using radix ten. That is, $f(10) = 8(10) + 9 = 89$. \subsection{Multiplication by $x$} Given a polynomial in $x$ such as $f(x) = a_n x^n + a_{n-1} x^{n-1} + ... + a_0$ multiplying by $x$ amounts to shifting the coefficients up one degree. In this case $f(x) \cdot x = a_n x^{n+1} + a_{n-1} x^n + ... + a_0 x$. From a scalar basis point of view multiplying by $x$ is equivalent to multiplying by the integer $\beta$. \newpage\begin{figure}[!here] \begin{small} \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{mp\_lshd}. \\ \textbf{Input}. One mp\_int $a$ and an integer $b$ \\ \textbf{Output}. $a \leftarrow a \cdot \beta^b$ (equivalent to multiplication by $x^b$). \\ \hline \\ 1. If $b \le 0$ then return(\textit{MP\_OKAY}). \\ 2. If $a.alloc < a.used + b$ then grow $a$ to at least $a.used + b$ digits. (\textit{mp\_grow}). \\ 3. If the reallocation failed return(\textit{MP\_MEM}). \\ 4. $a.used \leftarrow a.used + b$ \\ 5. $i \leftarrow a.used - 1$ \\ 6. $j \leftarrow a.used - 1 - b$ \\ 7. for $n$ from $a.used - 1$ to $b$ do \\ \hspace{3mm}7.1 $a_{i} \leftarrow a_{j}$ \\ \hspace{3mm}7.2 $i \leftarrow i - 1$ \\ \hspace{3mm}7.3 $j \leftarrow j - 1$ \\ 8. for $n$ from 0 to $b - 1$ do \\ \hspace{3mm}8.1 $a_n \leftarrow 0$ \\ 9. Return(\textit{MP\_OKAY}). \\ \hline \end{tabular} \end{center} \end{small} \caption{Algorithm mp\_lshd} \end{figure} \textbf{Algorithm mp\_lshd.} This algorithm multiplies an mp\_int by the $b$'th power of $x$. This is equivalent to multiplying by $\beta^b$. The algorithm differs from the other algorithms presented so far as it performs the operation in place instead storing the result in a separate location. The motivation behind this change is due to the way this function is typically used. Algorithms such as mp\_add store the result in an optionally different third mp\_int because the original inputs are often still required. Algorithm mp\_lshd (\textit{and similarly algorithm mp\_rshd}) is typically used on values where the original value is no longer required. The algorithm will return success immediately if $b \le 0$ since the rest of algorithm is only valid when $b > 0$. First the destination $a$ is grown as required to accomodate the result. The counters $i$ and $j$ are used to form a \textit{sliding window} over the digits of $a$ of length $b$. The head of the sliding window is at $i$ (\textit{the leading digit}) and the tail at $j$ (\textit{the trailing digit}). The loop on step 7 copies the digit from the tail to the head. In each iteration the window is moved down one digit. The last loop on step 8 sets the lower $b$ digits to zero. \newpage \begin{center} \begin{figure}[here] \includegraphics{pics/sliding_window.ps} \caption{Sliding Window Movement} \end{figure} \end{center} \index{bn\_mp\_lshd.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_mp\_lshd.c \vspace{-3mm} \begin{alltt} 016 017 /* shift left a certain amount of digits */ 018 int 019 mp_lshd (mp_int * a, int b) 020 \{ 021 int x, res; 022 023 /* if its less than zero return */ 024 if (b <= 0) \{ 025 return MP_OKAY; 026 \} 027 028 /* grow to fit the new digits */ 029 if (a->alloc < a->used + b) \{ 030 if ((res = mp_grow (a, a->used + b)) != MP_OKAY) \{ 031 return res; 032 \} 033 \} 034 035 \{ 036 register mp_digit *top, *bottom; 037 038 /* increment the used by the shift amount then copy upwards */ 039 a->used += b; 040 041 /* top */ 042 top = a->dp + a->used - 1; 043 044 /* base */ 045 bottom = a->dp + a->used - 1 - b; 046 047 /* much like mp_rshd this is implemented using a sliding window 048 * except the window goes the otherway around. Copying from 049 * the bottom to the top. see bn_mp_rshd.c for more info. 050 */ 051 for (x = a->used - 1; x >= b; x--) \{ 052 *top-- = *bottom--; 053 \} 054 055 /* zero the lower digits */ 056 top = a->dp; 057 for (x = 0; x < b; x++) \{ 058 *top++ = 0; 059 \} 060 \} 061 return MP_OKAY; 062 \} \end{alltt} \end{small} The if statement on line 24 ensures that the $b$ variable is greater than zero. The \textbf{used} count is incremented by $b$ before the copy loop begins. This elminates the need for an additional variable in the for loop. The variable $top$ on line 42 is an alias for the leading digit while $bottom$ on line 45 is an alias for the trailing edge. The aliases form a window of exactly $b$ digits over the input. \subsection{Division by $x$} Division by powers of $x$ is easily achieved by shifting the digits right and removing any that will end up to the right of the zero'th digit. \newpage\begin{figure}[!here] \begin{small} \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{mp\_rshd}. \\ \textbf{Input}. One mp\_int $a$ and an integer $b$ \\ \textbf{Output}. $a \leftarrow a / \beta^b$ (Divide by $x^b$). \\ \hline \\ 1. If $b \le 0$ then return. \\ 2. If $a.used \le b$ then do \\ \hspace{3mm}2.1 Zero $a$. (\textit{mp\_zero}). \\ \hspace{3mm}2.2 Return. \\ 3. $i \leftarrow 0$ \\ 4. $j \leftarrow b$ \\ 5. for $n$ from 0 to $a.used - b - 1$ do \\ \hspace{3mm}5.1 $a_i \leftarrow a_j$ \\ \hspace{3mm}5.2 $i \leftarrow i + 1$ \\ \hspace{3mm}5.3 $j \leftarrow j + 1$ \\ 6. for $n$ from $a.used - b$ to $a.used - 1$ do \\ \hspace{3mm}6.1 $a_n \leftarrow 0$ \\ 7. $a.used \leftarrow a.used - b$ \\ 8. Return. \\ \hline \end{tabular} \end{center} \end{small} \caption{Algorithm mp\_rshd} \end{figure} \textbf{Algorithm mp\_rshd.} This algorithm divides the input in place by the $b$'th power of $x$. It is analogous to dividing by a $\beta^b$ but much quicker since it does not require single precision division. This algorithm does not actually return an error code as it cannot fail. If the input $b$ is less than one the algorithm quickly returns without performing any work. If the \textbf{used} count is less than or equal to the shift count $b$ then it will simply zero the input and return. After the trivial cases of inputs have been handled the sliding window is setup. Much like the case of algorithm mp\_lshd a sliding window that is $b$ digits wide is used to copy the digits. Unlike mp\_lshd the window slides in the opposite direction from the trailing to the leading digit. Also the digits are copied from the leading to the trailing edge. Once the window copy is complete the upper digits must be zeroed and the \textbf{used} count decremented. \index{bn\_mp\_rshd.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_mp\_rshd.c \vspace{-3mm} \begin{alltt} 016 017 /* shift right a certain amount of digits */ 018 void 019 mp_rshd (mp_int * a, int b) 020 \{ 021 int x; 022 023 /* if b <= 0 then ignore it */ 024 if (b <= 0) \{ 025 return; 026 \} 027 028 /* if b > used then simply zero it and return */ 029 if (a->used <= b) \{ 030 mp_zero (a); 031 return; 032 \} 033 034 \{ 035 register mp_digit *bottom, *top; 036 037 /* shift the digits down */ 038 039 /* bottom */ 040 bottom = a->dp; 041 042 /* top [offset into digits] */ 043 top = a->dp + b; 044 045 /* this is implemented as a sliding window where 046 * the window is b-digits long and digits from 047 * the top of the window are copied to the bottom 048 * 049 * e.g. 050 051 b-2 | b-1 | b0 | b1 | b2 | ... | bb | ----> 052 /\symbol{92} | ----> 053 \symbol{92}-------------------/ ----> 054 */ 055 for (x = 0; x < (a->used - b); x++) \{ 056 *bottom++ = *top++; 057 \} 058 059 /* zero the top digits */ 060 for (; x < a->used; x++) \{ 061 *bottom++ = 0; 062 \} 063 \} 064 065 /* remove excess digits */ 066 a->used -= b; 067 \} \end{alltt} \end{small} The only noteworthy element of this routine is the lack of a return type. -- Will update later to give it a return type...Tom \section{Powers of Two} Now that algorithms for moving single bits as well as whole digits exist algorithms for moving the ``in between'' distances are required. For example, to quickly multiply by $2^k$ for any $k$ without using a full multiplier algorithm would prove useful. Instead of performing single shifts $k$ times to achieve a multiplication by $2^{\pm k}$ a mixture of whole digit shifting and partial digit shifting is employed. \subsection{Multiplication by Power of Two} \newpage\begin{figure}[!here] \begin{small} \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{mp\_mul\_2d}. \\ \textbf{Input}. One mp\_int $a$ and an integer $b$ \\ \textbf{Output}. $c \leftarrow a \cdot 2^b$. \\ \hline \\ 1. $c \leftarrow a$. (\textit{mp\_copy}) \\ 2. If $c.alloc < c.used + \lfloor b / lg(\beta) \rfloor + 2$ then grow $c$ accordingly. \\ 3. If the reallocation failed return(\textit{MP\_MEM}). \\ 4. If $b \ge lg(\beta)$ then \\ \hspace{3mm}4.1 $c \leftarrow c \cdot \beta^{\lfloor b / lg(\beta) \rfloor}$ (\textit{mp\_lshd}). \\ \hspace{3mm}4.2 If step 4.1 failed return(\textit{MP\_MEM}). \\ 5. $d \leftarrow b \mbox{ (mod }lg(\beta)\mbox{)}$ \\ 6. If $d \ne 0$ then do \\ \hspace{3mm}6.1 $mask \leftarrow 2^d$ \\ \hspace{3mm}6.2 $r \leftarrow 0$ \\ \hspace{3mm}6.3 for $n$ from $0$ to $c.used - 1$ do \\ \hspace{6mm}6.3.1 $rr \leftarrow c_n >> (lg(\beta) - d) \mbox{ (mod }mask\mbox{)}$ \\ \hspace{6mm}6.3.2 $c_n \leftarrow (c_n << d) + r \mbox{ (mod }\beta\mbox{)}$ \\ \hspace{6mm}6.3.3 $r \leftarrow rr$ \\ \hspace{3mm}6.4 If $r > 0$ then do \\ \hspace{6mm}6.4.1 $c_{c.used} \leftarrow r$ \\ \hspace{6mm}6.4.2 $c.used \leftarrow c.used + 1$ \\ 7. Return(\textit{MP\_OKAY}). \\ \hline \end{tabular} \end{center} \end{small} \caption{Algorithm mp\_mul\_2d} \end{figure} \textbf{Algorithm mp\_mul\_2d.} This algorithm multiplies $a$ by $2^b$ and stores the result in $c$. The algorithm uses algorithm mp\_lshd and a derivative of algorithm mp\_mul\_2 to quickly compute the product. First the algorithm will multiply $a$ by $x^{\lfloor b / lg(\beta) \rfloor}$ which will ensure that the remainder multiplicand is less than $\beta$. For example, if $b = 37$ and $\beta = 2^{28}$ then this step will multiply by $x$ leaving a multiplication by $2^{37 - 28} = 2^{9}$ left. After the digits have been shifted appropriately at most $lg(\beta) - 1$ shifts are left to perform. Step 5 calculates the number of remaining shifts required. If it is non-zero a modified shift loop is used to calculate the remaining product. Essentially the loop is a generic version of algorith mp\_mul2 designed to handle any shift count in the range $1 \le x < lg(\beta)$. The $mask$ variable is used to extract the upper $d$ bits to form the carry for the next iteration. This algorithm is loosely measured as a $O(2n)$ algorithm which means that if the input is $n$-digits that it takes $2n$ ``time'' to complete. It is possible to optimize this algorithm down to a $O(n)$ algorithm at a cost of making the algorithm slightly harder to follow. \index{bn\_mp\_mul\_2d.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_mp\_mul\_2d.c \vspace{-3mm} \begin{alltt} 016 017 /* NOTE: This routine requires updating. For instance the c->used = c->all oc bit 018 is wrong. We should just shift c->used digits then set the carry as c->d p[c->used] = carry 019 020 To be fixed for LTM 0.18 021 */ 022 023 /* shift left by a certain bit count */ 024 int 025 mp_mul_2d (mp_int * a, int b, mp_int * c) 026 \{ 027 mp_digit d; 028 int res; 029 030 /* copy */ 031 if (a != c) \{ 032 if ((res = mp_copy (a, c)) != MP_OKAY) \{ 033 return res; 034 \} 035 \} 036 037 if (c->alloc < (int)(c->used + b/DIGIT_BIT + 2)) \{ 038 if ((res = mp_grow (c, c->used + b / DIGIT_BIT + 2)) != MP_OKAY) \{ 039 return res; 040 \} 041 \} 042 043 /* shift by as many digits in the bit count */ 044 if (b >= (int)DIGIT_BIT) \{ 045 if ((res = mp_lshd (c, b / DIGIT_BIT)) != MP_OKAY) \{ 046 return res; 047 \} 048 \} 049 c->used = c->alloc; 050 051 /* shift any bit count < DIGIT_BIT */ 052 d = (mp_digit) (b % DIGIT_BIT); 053 if (d != 0) \{ 054 register mp_digit *tmpc, mask, r, rr; 055 register int x; 056 057 /* bitmask for carries */ 058 mask = (((mp_digit)1) << d) - 1; 059 060 /* alias */ 061 tmpc = c->dp; 062 063 /* carry */ 064 r = 0; 065 for (x = 0; x < c->used; x++) \{ 066 /* get the higher bits of the current word */ 067 rr = (*tmpc >> (DIGIT_BIT - d)) & mask; 068 069 /* shift the current word and OR in the carry */ 070 *tmpc = ((*tmpc << d) | r) & MP_MASK; 071 ++tmpc; 072 073 /* set the carry to the carry bits of the current word */ 074 r = rr; 075 \} 076 \} 077 mp_clamp (c); 078 return MP_OKAY; 079 \} \end{alltt} \end{small} Notes to be revised when code is updated. -- Tom \subsection{Division by Power of Two} \newpage\begin{figure}[!here] \begin{small} \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{mp\_div\_2d}. \\ \textbf{Input}. One mp\_int $a$ and an integer $b$ \\ \textbf{Output}. $c \leftarrow \lfloor a / 2^b \rfloor, d \leftarrow a \mbox{ (mod }2^b\mbox{)}$. \\ \hline \\ 1. If $b \le 0$ then do \\ \hspace{3mm}1.1 $c \leftarrow a$ (\textit{mp\_copy}) \\ \hspace{3mm}1.2 $d \leftarrow 0$ (\textit{mp\_zero}) \\ \hspace{3mm}1.3 Return(\textit{MP\_OKAY}). \\ 2. $c \leftarrow a$ \\ 3. $d \leftarrow a \mbox{ (mod }2^b\mbox{)}$ (\textit{mp\_mod\_2d}) \\ 4. If $b \ge lg(\beta)$ then do \\ \hspace{3mm}4.1 $c \leftarrow \lfloor c/\beta^{\lfloor b/lg(\beta) \rfloor} \rfloor$ (\textit{mp\_rshd}). \\ 5. $k \leftarrow b \mbox{ (mod }lg(\beta)\mbox{)}$ \\ 6. If $k \ne 0$ then do \\ \hspace{3mm}6.1 $mask \leftarrow 2^k$ \\ \hspace{3mm}6.2 $r \leftarrow 0$ \\ \hspace{3mm}6.3 for $n$ from $c.used - 1$ to $0$ do \\ \hspace{6mm}6.3.1 $rr \leftarrow c_n \mbox{ (mod }mask\mbox{)}$ \\ \hspace{6mm}6.3.2 $c_n \leftarrow (c_n >> k) + (r << (lg(\beta) - k))$ \\ \hspace{6mm}6.3.3 $r \leftarrow rr$ \\ 7. Clamp excess digits of $c$. (\textit{mp\_clamp}) \\ 8. Return(\textit{MP\_OKAY}). \\ \hline \end{tabular} \end{center} \end{small} \caption{Algorithm mp\_div\_2d} \end{figure} \textbf{Algorithm mp\_div\_2d.} This algorithm will divide an input $a$ by $2^b$ and produce the quotient and remainder. The algorithm is designed much like algorithm mp\_mul\_2d by first using whole digit shifts then single precision shifts. This algorithm will also produce the remainder of the division by using algorithm mp\_mod\_2d. \index{bn\_mp\_div\_2d.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_mp\_div\_2d.c \vspace{-3mm} \begin{alltt} 016 017 /* shift right by a certain bit count (store quotient in c, optional remaind er in d) */ 018 int 019 mp_div_2d (mp_int * a, int b, mp_int * c, mp_int * d) 020 \{ 021 mp_digit D, r, rr; 022 int x, res; 023 mp_int t; 024 025 026 /* if the shift count is <= 0 then we do no work */ 027 if (b <= 0) \{ 028 res = mp_copy (a, c); 029 if (d != NULL) \{ 030 mp_zero (d); 031 \} 032 return res; 033 \} 034 035 if ((res = mp_init (&t)) != MP_OKAY) \{ 036 return res; 037 \} 038 039 /* get the remainder */ 040 if (d != NULL) \{ 041 if ((res = mp_mod_2d (a, b, &t)) != MP_OKAY) \{ 042 mp_clear (&t); 043 return res; 044 \} 045 \} 046 047 /* copy */ 048 if ((res = mp_copy (a, c)) != MP_OKAY) \{ 049 mp_clear (&t); 050 return res; 051 \} 052 053 /* shift by as many digits in the bit count */ 054 if (b >= (int)DIGIT_BIT) \{ 055 mp_rshd (c, b / DIGIT_BIT); 056 \} 057 058 /* shift any bit count < DIGIT_BIT */ 059 D = (mp_digit) (b % DIGIT_BIT); 060 if (D != 0) \{ 061 register mp_digit *tmpc, mask; 062 063 /* mask */ 064 mask = (((mp_digit)1) << D) - 1; 065 066 /* alias */ 067 tmpc = c->dp + (c->used - 1); 068 069 /* carry */ 070 r = 0; 071 for (x = c->used - 1; x >= 0; x--) \{ 072 /* get the lower bits of this word in a temp */ 073 rr = *tmpc & mask; 074 075 /* shift the current word and mix in the carry bits from the previous word */ 076 *tmpc = (*tmpc >> D) | (r << (DIGIT_BIT - D)); 077 --tmpc; 078 079 /* set the carry to the carry bits of the current word found above */ 080 r = rr; 081 \} 082 \} 083 mp_clamp (c); 084 if (d != NULL) \{ 085 mp_exch (&t, d); 086 \} 087 mp_clear (&t); 088 return MP_OKAY; 089 \} \end{alltt} \end{small} The implementation of algorithm mp\_div\_2d is slightly different than the algorithm specifies. The remainder $d$ may be optionally ignored by passing \textbf{NULL} as the pointer to the mp\_int variable. The temporary mp\_int variable $t$ is used to hold the result of the remainder operation until the end. This allows $d$ and $a$ to represent the same mp\_int without modifying $a$ before the quotient is obtained. The remainder of the source code is essentially the same as the source code for mp\_mul\_2d. (-- Fix this paragraph up later, Tom). \subsection{Remainder of Division by Power of Two} The last algorithm in the series of polynomial basis power of two algorithms is calculating the remainder of division by $2^b$. This algorithm benefits from the fact that in twos complement arithmetic $a \mbox{ (mod }2^b\mbox{)}$ is the same as $a$ AND $2^b - 1$. \begin{figure}[!here] \begin{small} \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{mp\_mod\_2d}. \\ \textbf{Input}. One mp\_int $a$ and an integer $b$ \\ \textbf{Output}. $c \leftarrow a \mbox{ (mod }2^b\mbox{)}$. \\ \hline \\ 1. If $b \le 0$ then do \\ \hspace{3mm}1.1 $c \leftarrow 0$ (\textit{mp\_zero}) \\ \hspace{3mm}1.2 Return(\textit{MP\_OKAY}). \\ 2. If $b > a.used \cdot lg(\beta)$ then do \\ \hspace{3mm}2.1 $c \leftarrow a$ (\textit{mp\_copy}) \\ \hspace{3mm}2.2 Return the result of step 2.1. \\ 3. $c \leftarrow a$ \\ 4. If step 3 failed return(\textit{MP\_MEM}). \\ 5. for $n$ from $\lceil b / lg(\beta) \rceil$ to $c.used$ do \\ \hspace{3mm}5.1 $c_n \leftarrow 0$ \\ 6. $k \leftarrow b \mbox{ (mod }lg(\beta)\mbox{)}$ \\ 7. $c_{\lfloor b / lg(\beta) \rfloor} \leftarrow c_{\lfloor b / lg(\beta) \rfloor} \mbox{ (mod }2^{k}\mbox{)}$. \\ 8. Clamp excess digits of $c$. (\textit{mp\_clamp}) \\ 9. Return(\textit{MP\_OKAY}). \\ \hline \end{tabular} \end{center} \end{small} \caption{Algorithm mp\_mod\_2d} \end{figure} \textbf{Algorithm mp\_mod\_2d.} This algorithm will quickly calculate the value of $a \mbox{ (mod }2^b\mbox{)}$. First if $b$ is less than or equal to zero the result is set to zero. If $b$ is greater than the number of bits in $a$ then it simply copies $a$ to $c$ and returns. Otherwise, $a$ is copied to $b$, leading digits are removed and the remaining leading digit is trimed to the exact bit count. \index{bn\_mp\_mod\_2d.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_mp\_mod\_2d.c \vspace{-3mm} \begin{alltt} 016 017 /* calc a value mod 2\b */ 018 int 019 mp_mod_2d (mp_int * a, int b, mp_int * c) 020 \{ 021 int x, res; 022 023 024 /* if b is <= 0 then zero the int */ 025 if (b <= 0) \{ 026 mp_zero (c); 027 return MP_OKAY; 028 \} 029 030 /* if the modulus is larger than the value than return */ 031 if (b > (int) (a->used * DIGIT_BIT)) \{ 032 res = mp_copy (a, c); 033 return res; 034 \} 035 036 /* copy */ 037 if ((res = mp_copy (a, c)) != MP_OKAY) \{ 038 return res; 039 \} 040 041 /* zero digits above the last digit of the modulus */ 042 for (x = (b / DIGIT_BIT) + ((b % DIGIT_BIT) == 0 ? 0 : 1); x < c->used; x+ +) \{ 043 c->dp[x] = 0; 044 \} 045 /* clear the digit that is not completely outside/inside the modulus */ 046 c->dp[b / DIGIT_BIT] &= 047 (mp_digit) ((((mp_digit) 1) << (((mp_digit) b) % DIGIT_BIT)) - ((mp_digi t) 1)); 048 mp_clamp (c); 049 return MP_OKAY; 050 \} \end{alltt} \end{small} -- Add comments later, Tom. \section*{Exercises} \begin{tabular}{cl} $\left [ 3 \right ] $ & Devise an algorithm that performs $a \cdot 2^b$ for generic values of $b$ \\ & in $O(n)$ time. \\ &\\ $\left [ 3 \right ] $ & Devise an efficient algorithm to multiply by small low hamming \\ & weight values such as $3$, $5$ and $9$. Extend it to handle all values \\ & upto $64$ with a hamming weight less than three. \\ &\\ $\left [ 2 \right ] $ & Modify the preceding algorithm to handle values of the form \\ & $2^k - 1$ as well. \\ &\\ $\left [ 3 \right ] $ & Using only algorithms mp\_mul\_2, mp\_div\_2 and mp\_add create an \\ & algorithm to multiply two integers in roughly $O(2n^2)$ time for \\ & any $n$-bit input. Note that the time of addition is ignored in the \\ & calculation. \\ & \\ $\left [ 5 \right ] $ & Improve the previous algorithm to have a working time of at most \\ & $O \left (2^{(k-1)}n + \left ({2n^2 \over k} \right ) \right )$ for an appropriate choice of $k$. Again ignore \\ & the cost of addition. \\ & \\ $\left [ 2 \right ] $ & Devise a chart to find optimal values of $k$ for the previous problem \\ & for $n = 64 \ldots 1024$ in steps of $64$. \\ & \\ $\left [ 2 \right ] $ & Using only algorithms mp\_abs and mp\_sub devise another method for \\ & calculating the result of a signed comparison. \\ & \end{tabular} \chapter{Multiplication and Squaring} \section{The Multipliers} For most number theoretic systems including public key cryptographic algorithms the set of algorithms collectively known as the ``multipliers'' form the most important subset of algorithms of any multiple precision integer package. The set of multipliers include multiplication, squaring and modular reduction algorithms. The importance of these algorithms is driven by the fact that most popular public key algorithms are based on modular exponentiation. That is performing $d \equiv a^b \mbox{ (mod }c\mbox{)}$ for some arbitrary choice of $a$, $b$, $c$ and $d$. Roughly speaking the a modular exponentiation will spend about 40\% of the time in modular reductions, 35\% of the time in squaring and 25\% of the time in multiplications. Only a small trivial amount of time is spent on lower level algorithms such as mp\_clamp, mp\_init, etc... This chapter will discuss only two of the multipliers algorithms, multiplication and squaring. As will be discussed shortly very efficient multiplier algorithms are not always straightforward and deserve a lot of attention. \section{Multiplication} \subsection{The Baseline Multiplication} \index{baseline multiplication} Computing the product of two integers in software can be achieved using a trivial adaptation of the standard $O(n^2)$ long-hand multiplication algorithm school children are taught. The ``baseline multiplication'' algorithm is designed to act as the ``catch-all'' algorithm only called when the faster algorithms cannot be used. This algorithm does not use any particularly interesting optimizations. The first algorithm to review is the unsigned multiplication algorithm from which a signed multiplication algorithm can be established. One important facet of this algorithm to note is that it has been modified to only produce a certain amount of output digits as resolution. Recall that for a $n$ and $m$ digit input the product will be at most $n + m + 1$ digits. Therefore, this algorithm can be reduced to a full multiplier by telling it to produce $n + m + 1$ digits. Recall from sub-section 5.2.2 the definition of $\gamma$ as the number of bits in the type \textbf{mp\_digit}. We shall now extend this variable set to include $\alpha$ which shall represent the number of bits in the type \textbf{mp\_word}. This implies that $2^{\alpha} > 2 \cdot \beta^2$. The constant $\delta = 2^{\alpha - 2lg(\beta)}$ will represent the maximal weight of any column in a product (\textit{see sub-section 6.2.2 for more information}). \newpage\begin{figure}[!here] \begin{small} \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{s\_mp\_mul\_digs}. \\ \textbf{Input}. mp\_int $a$, mp\_int $b$ and an integer $digs$ \\ \textbf{Output}. $c \leftarrow \vert a \vert \cdot \vert b \vert \mbox{ (mod }\beta^{digs}\mbox{)}$. \\ \hline \\ 1. If min$(a.used, b.used) < \delta$ then do \\ \hspace{3mm}1.1 Calculate $c = \vert a \vert \cdot \vert b \vert$ by the Comba method. \\ \hspace{3mm}1.2 Return the result of step 1.1 \\ \\ Allocate and initialize a temporary mp\_int. \\ 2. Init $t$ to be of size $digs$ \\ 3. If step 2 failed return(\textit{MP\_MEM}). \\ 4. $t.used \leftarrow digs$ \\ \\ Compute the product. \\ 5. for $ix$ from $0$ to $a.used - 1$ do \\ \hspace{3mm}5.1 $u \leftarrow 0$ \\ \hspace{3mm}5.2 $pb \leftarrow \mbox{min}(b.used, digs - ix)$ \\ \hspace{3mm}5.3 If $pb < 1$ then goto step 6. \\ \hspace{3mm}5.4 for $iy$ from $0$ to $pb - 1$ do \\ \hspace{6mm}5.4.1 $\hat r \leftarrow t_{iy + ix} + a_{ix} \cdot b_{iy} + u$ \\ \hspace{6mm}5.4.2 $t_{iy + ix} \leftarrow \hat r \mbox{ (mod }\beta\mbox{)}$ \\ \hspace{6mm}5.4.3 $u \leftarrow \lfloor \hat r / \beta \rfloor$ \\ \hspace{3mm}5.5 if $ix + iy < digs$ then do \\ \hspace{6mm}5.5.1 $t_{ix + pb} \leftarrow u$ \\ 6. Clamp excess digits of $t$. \\ 7. Swap $c$ with $t$ \\ 8. Clear $t$ \\ 9. Return(\textit{MP\_OKAY}). \\ \hline \end{tabular} \end{center} \end{small} \caption{Algorithm s\_mp\_mul\_digs} \end{figure} \textbf{Algorithm s\_mp\_mul\_digs.} This algorithm computes the unsigned product of two inputs $a$ and $b$ limited to an output precision of $digs$ digits. While it may seem a bit awkward to modify the function from its simple $O(n^2)$ description the usefulness of partial multipliers will arise in a future algorithm. The algorithm is loosely based on algorithm 14.12 from \cite[pp. 595]{HAC} and is similar to Algorithm M \cite[pp. 268]{TAOCPV2}. The algorithm differs from those cited references because it can produce a variable output precision regardless of the precision of the inputs. The first thing this algorithm checks for is whether a Comba multiplier can be used instead. That is if the minimal digit count of either input is less than $\delta$ the Comba method is used. After the Comba method is ruled out the baseline algorithm begins. A temporary mp\_int variable $t$ is used to hold the intermediate result of the product. This allows the algorithm to be used to compute products when either $a = c$ or $b = c$ without overwriting the inputs. All of step 5 is the infamous $O(n^2)$ multiplication loop slightly modified to only produce upto $digs$ digits of output. The $pb$ variable is given the count of digits to read from $b$ inside the nested loop. If $pb < 0$ then no more output digits can be produced and the algorithm will exit the loop. The best way to think of the loops are as a series of $pb \times 1$ multiplication. That is, in each pass of the innermost loop $a_{ix}$ is multiplied against $b$ and the result is added (\textit{with an appropriate shift}) to $t$. For example, consider multiplying $576$ by $241$. That is equivalent to computing $10^0(1)(576) + 10^1(4)(576) + 10^2(2)(576)$ which is best visualized as the following table. \begin{figure}[here] \begin{center} \begin{tabular}{|c|c|c|c|c|c|c|} \hline && & 5 & 7 & 6 & \\ \hline $\times$&& & 2 & 4 & 1 & \\ \hline &&&&&&\\ && & 5 & 7 & 6 & $10^0(1)(576)$ \\ &2 & 3 & 0 & 4 & 0 & $10^1(4)(576)$ \\ 1 & 1 & 5 & 2 & 0 & 0 & $10^2(2)(576)$ \\ \hline \end{tabular} \end{center} \caption{Long-Hand Multiplication Diagram} \end{figure} Each row of the product is added to the result after being shifted to the left (\textit{multiplied by a power of the radix}) by the appropriate count. That is in pass $ix$ of the inner loop the product is added starting at the $ix$'th digit of the reult. Step 5.4.1 introduces the hat symbol (\textit{e.g. $\hat x$}) which represents a double precision variable. The multiplication on that step is assumed to be a double wide output single precision multiplication. That is, two single precision variables are multiplied to produce a double precision result. The step is somewhat optimized from a long-hand multiplication algorithm because the carry from the addition in step 5.4.1 is forwarded through the nested loop. If the carry was ignored it would overflow the single precision digit $t_{ix+iy}$ and the result would be lost. At step 5.5 the nested loop is finished and any carry that was left over should be forwarded. That is provided $ix + iy < digs$ otherwise the carry is ignored since it will not be part of the result anyways. \index{bn\_s\_mp\_mul\_digs.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_s\_mp\_mul\_digs.c \vspace{-3mm} \begin{alltt} 016 017 /* multiplies |a| * |b| and only computes upto digs digits of result 018 * HAC pp. 595, Algorithm 14.12 Modified so you can control how 019 * many digits of output are created. 020 */ 021 int 022 s_mp_mul_digs (mp_int * a, mp_int * b, mp_int * c, int digs) 023 \{ 024 mp_int t; 025 int res, pa, pb, ix, iy; 026 mp_digit u; 027 mp_word r; 028 mp_digit tmpx, *tmpt, *tmpy; 029 030 /* can we use the fast multiplier? */ 031 if (((digs) < MP_WARRAY) && 032 MIN (a->used, b->used) < 033 (1 << ((CHAR_BIT * sizeof (mp_word)) - (2 * DIGIT_BIT)))) \{ 034 return fast_s_mp_mul_digs (a, b, c, digs); 035 \} 036 037 if ((res = mp_init_size (&t, digs)) != MP_OKAY) \{ 038 return res; 039 \} 040 t.used = digs; 041 042 /* compute the digits of the product directly */ 043 pa = a->used; 044 for (ix = 0; ix < pa; ix++) \{ 045 /* set the carry to zero */ 046 u = 0; 047 048 /* limit ourselves to making digs digits of output */ 049 pb = MIN (b->used, digs - ix); 050 051 /* setup some aliases */ 052 /* copy of the digit from a used within the nested loop */ 053 tmpx = a->dp[ix]; 054 055 /* an alias for the destination shifted ix places */ 056 tmpt = t.dp + ix; 057 058 /* an alias for the digits of b */ 059 tmpy = b->dp; 060 061 /* compute the columns of the output and propagate the carry */ 062 for (iy = 0; iy < pb; iy++) \{ 063 /* compute the column as a mp_word */ 064 r = ((mp_word) *tmpt) + 065 ((mp_word) tmpx) * ((mp_word) * tmpy++) + 066 ((mp_word) u); 067 068 /* the new column is the lower part of the result */ 069 *tmpt++ = (mp_digit) (r & ((mp_word) MP_MASK)); 070 071 /* get the carry word from the result */ 072 u = (mp_digit) (r >> ((mp_word) DIGIT_BIT)); 073 \} 074 /* set carry if it is placed below digs */ 075 if (ix + iy < digs) \{ 076 *tmpt = u; 077 \} 078 \} 079 080 mp_clamp (&t); 081 mp_exch (&t, c); 082 083 mp_clear (&t); 084 return MP_OKAY; 085 \} \end{alltt} \end{small} Lines 31 to 35 determine if the Comba method can be used first. The conditions for using the Comba routine are that min$(a.used, b.used) < \delta$ and the number of digits of output is less than \textbf{MP\_WARRAY}. This new constant is used to control the stack usage in the Comba routines. By default it is set to $\delta$ but can be reduced when memory is at a premium. Of particular importance is the calculation of the $ix+iy$'th column on lines 64, 65 and 66. Note how all of the variables are cast to the type \textbf{mp\_word}. That is to ensure that double precision operations are used instead of single precision. The multiplication on line 65 is a bit of a GCC optimization. On the outset it looks like the compiler will have to use a double precision multiplication to produce the result required. Such an operation would be horribly slow on most processors and drag this to a crawl. However, GCC is smart enough to realize that double wide output single precision multipliers can be used. For example, the instruction ``MUL'' on the x86 processor can multiply two 32-bit values and produce a 64-bit result. \subsection{Faster Multiplication by the ``Comba'' Method} One of the huge drawbacks of the ``baseline'' algorithms is that at the $O(n^2)$ level the carry must be computed and propagated upwards. This makes the nested loop very sequential and hard to unroll and implement in parallel. The ``Comba'' method is named after little known (\textit{in cryptographic venues}) Paul G. Comba where in \cite{COMBA} a method of implementing fast multipliers that do not require nested carry fixup operations was presented. As an interesting aside it seems that Paul Barrett describes a similar technique in his 1986 paper \cite{BARRETT} which was written five years before \cite{COMBA}. At the heart of algorithm is once again the long-hand algorithm for multiplication. Except in this case a slight twist is placed on how the columns of the result are produced. In the standard long-hand algorithm rows of products are produced then added together to form the final result. In the baseline algorithm the columns are added together to get the result instantaneously. In the Comba algorithm however, the columns of the result are produced entirely independently of each other. That is at the $O(n^2)$ level a simple multiplication and addition step is performed. Or more succintly that \begin{equation} x_n = \sum_{i+j = n} a_ib_j \end{equation} Where $x_n$ is the $n'th$ column of the output vector. To see how this works consider once again multiplying $576$ by $241$. \begin{figure}[here] \begin{small} \begin{center} \begin{tabular}{|c|c|c|c|c|c|} \hline & & 5 & 7 & 6 & First Input\\ \hline $\times$ & & 2 & 4 & 1 & Second Input\\ \hline & & $1 \cdot 5 = 5$ & $1 \cdot 7 = 7$ & $1 \cdot 6 = 6$ & First pass \\ & $4 \cdot 5 = 20$ & $4 \cdot 7+5=33$ & $4 \cdot 6+7=31$ & 6 & Second pass \\ $2 \cdot 5 = 10$ & $2 \cdot 7 + 20 = 34$ & $2 \cdot 6+33=45$ & 31 & 6 & Third pass \\ \hline 10 & 34 & 45 & 31 & 6 & Final Result \\ \hline \end{tabular} \end{center} \end{small} \caption{Comba Multiplication Diagram} \end{figure} At this point the vector $x = \left < 10, 34, 45, 31, 6 \right >$ is the result of the first step of the Comba multipler. Now the columns must be fixed by propagating the carry upwards. The following trivial algorithm will accomplish this. \begin{enumerate} \item for $n$ from 0 to $k - 1$ do \item \hspace{3mm} $x_{n+1} \leftarrow x_{n+1} + \lfloor x_{n}/\beta \rfloor$ \item \hspace{3mm} $x_{n} \leftarrow x_{n} \mbox{ (mod }\beta\mbox{)}$ \end{enumerate} With that algorithm and $k = 5$ and $\beta = 10$ the following vector is produced $y = \left < 1, 3, 8, 8, 1, 6 \right >$. In this case $241 \cdot 576$ is in fact $138816$ and the procedure succeeded. If the algorithm is correct and as will be demonstrated shortly more efficient than the baseline algorithm why not simply always use this algorithm? \subsubsection{Column Weight.} At the nested $O(n^2)$ level the Comba method adds the product of two single precision variables to a each column of the output independently. A serious obstacle is if the carry is lost due to lack of precision before the algorithm has a chance to fix the carries. For example, in the multiplication of two three-digit numbers the third column of output will be the sum of three single precision multiplications. If the precision of the accumulator for the output digits is less then $3 \cdot (\beta - 1)^2$ then an overflow can occur and the carry information will be lost. For any $m$ and $n$ digit input the maximal weight of any column is min$(m, n)$ which is fairly obvious. The maximal number of terms in any column of a product is known as the ``column weight'' and strictly governs when the algorithm can be used. Recall from earlier that a double precision type has $\alpha$ bits of resolution and a single precision digit has $lg(\beta)$ bits of precision. Given these two quantities we may not violate the following \begin{equation} k \cdot \left (\beta - 1 \right )^2 < 2^{\alpha} \end{equation} Which reduces to \begin{equation} k \cdot \left ( \beta^2 - 2\beta + 1 \right ) < 2^{\alpha} \end{equation} Let $\rho = lg(\beta)$ represent the number of bits in a single precision digit. By further re-arrangement of the equation the final solution is found. \begin{equation} k \cdot \left (2^{2\rho} - 2^{\rho + 1} + 1 \right ) < 2^{\alpha} \end{equation} The defaults for LibTomMath are $\beta = 2^{28}, \alpha = 2^{64}$ which simplies to $72057593501057025 \cdot k < 2^{64}$ which when divided out result in $k < 257$. This implies that the smallest input may not have more than $256$ digits if the Comba method is to be used in this configuration. This is quite satisfactory for most applications since $256$ digits would be allow for numbers in the range of $2^{7168}$ which is much larger than the typical $2^{100}$ to $2^{4000}$ range most public key cryptographic algorithms use. \newpage\begin{figure}[!here] \begin{small} \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{fast\_s\_mp\_mul\_digs}. \\ \textbf{Input}. mp\_int $a$, mp\_int $b$ and an integer $digs$ \\ \textbf{Output}. $c \leftarrow \vert a \vert \cdot \vert b \vert \mbox{ (mod }\beta^{digs}\mbox{)}$. \\ \hline \\ Place an array of \textbf{MP\_WARRAY} double precision digits named $\hat W$ on the stack. \\ 1. If $c.alloc < digs$ then grow $c$ to $digs$ digits. (\textit{mp\_grow}) \\ 2. If step 1 failed return(\textit{MP\_MEM}).\\ \\ Zero the temporary array $\hat W$. \\ 3. for $n$ from $0$ to $digs - 1$ do \\ \hspace{3mm}3.1 $\hat W_n \leftarrow 0$ \\ \\ Compute the columns. \\ 4. for $ix$ from $0$ to $a.used - 1$ do \\ \hspace{3mm}4.1 $pb \leftarrow \mbox{min}(b.used, digs - ix)$ \\ \hspace{3mm}4.2 If $pb < 1$ then goto step 5. \\ \hspace{3mm}4.3 for $iy$ from $0$ to $pb - 1$ do \\ \hspace{6mm}4.3.1 $\hat W_{ix+iy} \leftarrow \hat W_{ix+iy} + a_{ix}b_{iy}$ \\ \\ Propagate the carries upwards. \\ 5. $oldused \leftarrow c.used$ \\ 6. $c.used \leftarrow digs$ \\ 7. If $digs > 1$ then do \\ \hspace{3mm}7.1. for $ix$ from $1$ to $digs - 1$ do \\ \hspace{6mm}7.1.1 $\hat W_{ix} \leftarrow \hat W_{ix} + \lfloor \hat W_{ix-1} / \beta \rfloor$ \\ \hspace{6mm}7.1.2 $c_{ix - 1} \leftarrow \hat W_{ix - 1} \mbox{ (mod }\beta\mbox{)}$ \\ 8. else do \\ \hspace{3mm}8.1 $ix \leftarrow 0$ \\ 9. $c_{ix} \leftarrow \hat W_{ix} \mbox{ (mod }\beta\mbox{)}$ \\ \\ Zero excess digits. \\ 10. If $digs < oldused$ then do \\ \hspace{3mm}10.1 for $n$ from $digs$ to $oldused - 1$ do \\ \hspace{6mm}10.1.1 $c_n \leftarrow 0$ \\ 11. Clamp excessive digits of $c$. (\textit{mp\_clamp}) \\ 12. Return(\textit{MP\_OKAY}). \\ \hline \end{tabular} \end{center} \end{small} \caption{Algorithm fast\_s\_mp\_mul\_digs} \end{figure} \textbf{Algorithm fast\_s\_mp\_mul\_digs.} This algorithm performs the unsigned multiplication of $a$ and $b$ using the Comba method limited to $digs$ digits of precision. The algorithm essentially peforms the same calculation as algorithm s\_mp\_mul\_digs but much faster. The array $\hat W$ is meant to be on the stack when the algorithm is used. The size of the array does not change which is ideal. Note also that unlike algorithm s\_mp\_mul\_digs no temporary mp\_int is required since the result is calculated in place in $\hat W$. The $O(n^2)$ loop on step four is where the Comba method starts to show through. First there is no carry variable in the loop. Second the double precision multiply and add step does not have a carry fixup of any sort. In fact the nested loop is very simple and can be implemented in parallel. What makes the Comba method so attractive is that the carry propagation only takes place outside the $O(n^2)$ nested loop. For example, if the cost in terms of time of a multiply and add is $p$ and the cost of a carry propagation is $q$ then a baseline multiplication would require $O \left ((p + q)n^2 \right )$ time to multiply two $n$-digit numbers. The Comba method only requires $pn^2 + qn$ time, however, in practice the speed increase is actually much more. With $O(n)$ space the algorithm can be reduced to $O(pn + qn)$ time by implementing the $n$ multiply and add operations in the nested loop in parallel. The carry propagation loop on step 7 is fairly straightforward. It could have been written phased the other direction, that is, to assign to $c_{ix}$ instead of $c_{ix-1}$ in each iteration. However, it would still require pre-caution to make sure that $\hat W_{ix+1}$ is not beyond the \textbf{MP\_WARRAY} words set aside. \index{bn\_fast\_s\_mp\_mul\_digs.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_fast\_s\_mp\_mul\_digs.c \vspace{-3mm} \begin{alltt} 016 017 /* Fast (comba) multiplier 018 * 019 * This is the fast column-array [comba] multiplier. It is 020 * designed to compute the columns of the product first 021 * then handle the carries afterwards. This has the effect 022 * of making the nested loops that compute the columns very 023 * simple and schedulable on super-scalar processors. 024 * 025 * This has been modified to produce a variable number of 026 * digits of output so if say only a half-product is required 027 * you don't have to compute the upper half (a feature 028 * required for fast Barrett reduction). 029 * 030 * Based on Algorithm 14.12 on pp.595 of HAC. 031 * 032 */ 033 int 034 fast_s_mp_mul_digs (mp_int * a, mp_int * b, mp_int * c, int digs) 035 \{ 036 int olduse, res, pa, ix; 037 mp_word W[MP_WARRAY]; 038 039 /* grow the destination as required */ 040 if (c->alloc < digs) \{ 041 if ((res = mp_grow (c, digs)) != MP_OKAY) \{ 042 return res; 043 \} 044 \} 045 046 /* clear temp buf (the columns) */ 047 memset (W, 0, sizeof (mp_word) * digs); 048 049 /* calculate the columns */ 050 pa = a->used; 051 for (ix = 0; ix < pa; ix++) \{ 052 /* this multiplier has been modified to allow you to 053 * control how many digits of output are produced. 054 * So at most we want to make upto "digs" digits of output. 055 * 056 * this adds products to distinct columns (at ix+iy) of W 057 * note that each step through the loop is not dependent on 058 * the previous which means the compiler can easily unroll 059 * the loop without scheduling problems 060 */ 061 \{ 062 register mp_digit tmpx, *tmpy; 063 register mp_word *_W; 064 register int iy, pb; 065 066 /* alias for the the word on the left e.g. A[ix] * A[iy] */ 067 tmpx = a->dp[ix]; 068 069 /* alias for the right side */ 070 tmpy = b->dp; 071 072 /* alias for the columns, each step through the loop adds a new 073 term to each column 074 */ 075 _W = W + ix; 076 077 /* the number of digits is limited by their placement. E.g. 078 we avoid multiplying digits that will end up above the # of 079 digits of precision requested 080 */ 081 pb = MIN (b->used, digs - ix); 082 083 for (iy = 0; iy < pb; iy++) \{ 084 *_W++ += ((mp_word) tmpx) * ((mp_word) * tmpy++); 085 \} 086 \} 087 088 \} 089 090 /* setup dest */ 091 olduse = c->used; 092 c->used = digs; 093 094 \{ 095 register mp_digit *tmpc; 096 097 /* At this point W[] contains the sums of each column. To get the 098 * correct result we must take the extra bits from each column and 099 * carry them down 100 * 101 * Note that while this adds extra code to the multiplier it 102 * saves time since the carry propagation is removed from the 103 * above nested loop.This has the effect of reducing the work 104 * from N*(N+N*c)==N**2 + c*N**2 to N**2 + N*c where c is the 105 * cost of the shifting. On very small numbers this is slower 106 * but on most cryptographic size numbers it is faster. 107 */ 108 tmpc = c->dp; 109 for (ix = 1; ix < digs; ix++) \{ 110 W[ix] += (W[ix - 1] >> ((mp_word) DIGIT_BIT)); 111 *tmpc++ = (mp_digit) (W[ix - 1] & ((mp_word) MP_MASK)); 112 \} 113 *tmpc++ = (mp_digit) (W[digs - 1] & ((mp_word) MP_MASK)); 114 115 /* clear unused */ 116 for (; ix < olduse; ix++) \{ 117 *tmpc++ = 0; 118 \} 119 \} 120 121 mp_clamp (c); 122 return MP_OKAY; 123 \} \end{alltt} \end{small} The memset on line 47 clears the initial $\hat W$ array to zero in a single step. Like the slower baseline multiplication implementation a series of aliases (\textit{lines 67, 70 and 75}) are used to simplify the inner $O(n^2)$ loop. In this case a new alias $\_\hat W$ has been added which refers to the double precision columns offset by $ix$ in each pass. The inner loop on line 84 is where the algorithm will spend the majority of the time. Which is why it has been stripped to the bones of any extra baggage\footnote{Hence the pointer aliases.}. On x86 processors the multiply and add amounts to at the very least five instructions (\textit{two loads, two additions, one multiply}) while on the ARMv4 processors it amounts to only three (\textit{one load, one store, one multiply-add}). On both the x86 and ARMv4 processors GCC v3.2 does a very good job at unrolling the loop and scheduling it so there are very few dependency stalls. In theory the difference between the baseline and comba algorithms is a mere $O(qn)$ time difference. However, in the $O(n^2)$ nested loop of the baseline method there are dependency stalls as the algorithm must wait for the multiplier to finish before propagating the carry to the next digit. As a result fewer of the often multiple execution units\footnote{The AMD Athlon has three execution units and the Intel P4 has four.} can be simultaneously used. \subsection{Polynomial Basis Multiplication} To break the $O(n^2)$ barrier in multiplication requires a completely different look at integer multiplication. In the following algorithms the use of polynomial basis representation for two integers $a$ and $b$ as $f(x) = \sum_{i=0}^{n} a_i x^i$ and $g(x) = \sum_{i=0}^{n} b_i x^i$. respectively, is required. In this system both $f(x)$ and $g(x)$ have $n + 1$ terms and are of the $n$'th degree. The product $a \cdot b \equiv f(x) \cdot g(x)$ is the polynomial $W(x) = \sum_{i=0}^{2n} w_i x^i$. The coefficients $w_i$ will directly yield the desired product when $\beta$ is substituted for $x$. The direct solution to solve for the $2n + 1$ coefficients requires $O(n^2)$ time and is would be in practice slower than the Comba technique. However, numerical analysis theory will indicate that only $2n + 1$ points in $W(x)$ are required to provide $2n + 1$ knowns for the $2n + 1$ unknowns. This means by finding $\zeta_y = W(y)$ for $2n + 1$ small values of $y$ the coefficients of $W(x)$ can be found with trivial Gaussian elimination. Since the polynomial $W(x)$ is unknown the equivalent $\zeta_y = f(y) \cdot g(y)$ is used in its place. The benefit of this technique stems from the fact that $f(y)$ and $g(y)$ are much smaller than either $a$ or $b$ respectively. In fact if both polynomials have $n + 1$ terms then the multiplicands will be $n$ times smaller than the inputs. Even if $2n + 1$ multiplications are required since they are of smaller values the algorithm is still faster. When picking points to gather relations there are always three obvious points to choose, $y = 0, 1$ and $ \infty$. The $\zeta_0$ term is simply the product $W(0) = w_0 = a_0 \cdot b_0$. The $\zeta_1$ term is the product $W(1) = \left (\sum_{i = 0}^{n} a_i \right ) \left (\sum_{i = 0}^{n} b_i \right )$. The third point $\zeta_{\infty}$ is less obvious but rather simple to explain. The $2n + 1$'th coefficient of $W(x)$ is numerically equivalent to the most significant column in an integer multiplication. The point at $\infty$ is used symbolically to represent the most significant column, that is $W(\infty) = w_{2n + 1} = a_nb_n$. Note that the points at $y = 0$ and $\infty$ yield the coefficients $w_0$ and $w_{2n + 1}$ directly. If more points are required they should be of small input values which are powers of two such as $2^q$ and the related \textit{mirror points} $\left (2^q \right )^{2n} \cdot \zeta_{2^{-q}}$ for small values of $q$. Using such points will allow the values of $f(y)$ and $g(y)$ to be independently calculated using only left shifts. As a general rule of the algorithm when the inputs are split into $n$ parts each there are $2n - 1$ multiplications. Each multiplication is of multiplicands that have $n$ times fewer digits than the inputs. The asymptotic running time of this algorithm is $O \left ( k^{lg_n(2n - 1)} \right )$ for $k$ digit inputs (\textit{assuming they have the same number of digits}). The following table summarizes the exponents for various values of $n$. \newpage\begin{figure} \begin{center} \begin{tabular}{|c|c|c|} \hline \textbf{Split into $n$ Parts} & \textbf{Exponent} & \textbf{Notes}\\ \hline $2$ & $1.584962501$ & This is Karatsuba Multiplication. \\ \hline $3$ & $1.464973520$ & This is Toom-Cook Multiplication. \\ \hline $4$ & $1.403677461$ &\\ \hline $5$ & $1.365212389$ &\\ \hline $10$ & $1.278753601$ &\\ \hline $100$ & $1.149426538$ &\\ \hline $1000$ & $1.100270931$ &\\ \hline $10000$ & $1.075252070$ &\\ \hline \end{tabular} \end{center} \caption{Asymptotic Running Time of Polynomial Basis Multiplication} \end{figure} At first it may seem like a good idea to choose $n = 1000$ since afterall the exponent is approximately $1.1$. However, the overhead of solving for the 2001 terms of $W(x)$ will certainly consume any savings the algorithm could offer for all but exceedingly large numbers. \subsubsection{Cutoff Point} The polynomial basis multiplication algorithms all require fewer single precision multiplications than a straight Comba approach. However, the algorithms incur an overhead (\textit{at the $O(n)$ work level}) since they require a system of equations to be solved. This makes them costly to use with small inputs. Let $m$ represent the number of digits in the multiplicands (\textit{assume both multiplicands have the same number of digits}). There exists a point $y$ such that when $m < y$ the polynomial basis algorithms are more costly than Comba, when $m = y$ they are roughly the same cost and when $m > y$ the Comba methods are slower than the polynomial basis algorithms. The exact location of $y$ depends on several key architectural elements of the computer platform in question. \begin{enumerate} \item The ratio of clock cycles for single precision multiplication versus other simpler operations such as addition, shifting, etc. For example on the AMD Athlon the ratio is roughly $17 : 1$ while on the Intel P4 it is $29 : 1$. The higher the ratio in favour of multiplication the lower the cutoff point $y$ will be. \item The complexity of the linear system of equations (\textit{for the coefficients of $W(x)$}) is. Generally speaking as the number of splits grows the complexity grows substantially. Ideally solving the system will only involve addition, subtraction and shifting of integers. This directly reflects on the ratio previous mentioned. \item To a lesser extent memory bandwidth and function call overheads. Provided the values are in the processor cache this is less of an influence over the cutoff point. \end{enumerate} A clean cutoff point separation occurs when a point $y$ is found such that all of the cutoff point conditions are met. For example, if the point is too low then there will be values of $m$ such that $m > y$ and the Comba method is still faster. Finding the cutoff points is fairly simple when a high resolution timer is available. \subsection{Karatsuba Multiplication} Karatsuba multiplication \cite{KARA} when originally proposed in 1962 was among the first set of algorithms to break the $O(n^2)$ barrier for general purpose multiplication. Given two polynomial basis representations $f(x) = ax + b$ and $g(x) = cx + d$ Karatsuba proved with light number theory \cite{KARAP} that the following polynomial is equivalent to multiplication of the two integers the polynomials represent. \begin{equation} f(x) \cdot g(x) = acx^2 + ((a - b)(c - d) + ac + bd)x + bd \end{equation} Using the observation that $ac$ and $bd$ could be re-used only three half sized multiplications would be required to produce the product. Applying this recursively the work factor becomes $O(n^{lg(3)})$ which is substantially better than the work factor $O(n^2)$ of the Comba technique. It turns out what Karatsuba did not know or at least did not publish was that this is simply polynomial basis multiplication with the points $\zeta_0$, $\zeta_{\infty}$ and $-\zeta_{-1}$. Consider the resultant system of equations. \begin{center} \begin{tabular}{rcrcrcrc} $\zeta_{0}$ & $=$ & & & & & $w_0$ \\ $-\zeta_{-1}$ & $=$ & $-w_2$ & $+$ & $w_1$ & $-$ & $w_0$ \\ $\zeta_{\infty}$ & $=$ & $w_2$ & & & & \\ \end{tabular} \end{center} By adding the first and last equation to the equation in the middle the term $w_1$ can be isolated and all three coefficients solved for. The simplicity of this system of equations has made Karatsuba fairly popular. In fact the cutoff point is often fairly low\footnote{With LibTomMath 0.18 it is 70 and 109 for the Intel P4 and AMD Athlon respectively.} making it an ideal algorithm to speed up certain public key cryptosystems such as RSA and Diffie-Hellman. It is worth noting that the point $\zeta_1$ could be substituted for $-\zeta_{-1}$. In this case the first and third row are subtracted instead of added to the second row. \newpage\begin{figure}[!here] \begin{small} \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{mp\_karatsuba\_mul}. \\ \textbf{Input}. mp\_int $a$ and mp\_int $b$ \\ \textbf{Output}. $c \leftarrow \vert a \vert \cdot \vert b \vert$ \\ \hline \\ 1. Init the following mp\_int variables: $x0$, $x1$, $y0$, $y1$, $t1$, $x0y0$, $x1y1$.\\ 2. If step 2 failed then return(\textit{MP\_MEM}). \\ \\ Split the input. e.g. $a = x1 \cdot \beta^B + x0$ \\ 3. $B \leftarrow \mbox{min}(a.used, b.used)/2$ \\ 4. $x0 \leftarrow a \mbox{ (mod }\beta^B\mbox{)}$ (\textit{mp\_mod\_2d}) \\ 5. $y0 \leftarrow b \mbox{ (mod }\beta^B\mbox{)}$ \\ 6. $x1 \leftarrow \lfloor a / \beta^B \rfloor$ (\textit{mp\_rshd}) \\ 7. $y1 \leftarrow \lfloor b / \beta^B \rfloor$ \\ \\ Calculate the three products. \\ 8. $x0y0 \leftarrow x0 \cdot y0$ (\textit{mp\_mul}) \\ 9. $x1y1 \leftarrow x1 \cdot y1$ \\ 10. $t1 \leftarrow x1 - x0$ (\textit{mp\_sub}) \\ 11. $x0 \leftarrow y1 - y0$ \\ 12. $t1 \leftarrow t1 \cdot x0$ \\ \\ Calculate the middle term. \\ 13. $x0 \leftarrow x0y0 + x1y1$ \\ 14. $t1 \leftarrow x0 - t1$ \\ \\ Calculate the final product. \\ 15. $t1 \leftarrow t1 \cdot \beta^B$ (\textit{mp\_lshd}) \\ 16. $x1y1 \leftarrow x1y1 \cdot \beta^{2B}$ \\ 17. $t1 \leftarrow x0y0 + t1$ \\ 18. $c \leftarrow t1 + x1y1$ \\ 19. Clear all of the temporary variables. \\ 20. Return(\textit{MP\_OKAY}).\\ \hline \end{tabular} \end{center} \end{small} \caption{Algorithm mp\_karatsuba\_mul} \end{figure} \textbf{Algorithm mp\_karatsuba\_mul.} This algorithm computes the unsigned product of two inputs using the Karatsuba method. It is loosely based on the description from \cite[pp. 294-295]{TAOCPV2}. \index{radix point} In order to split the two inputs into their respective halves a suitable \textit{radix point} must be chosen. The radix point chosen must be used for both of the inputs meaning that it must smaller than the smallest input. Step 3 chooses the radix point $B$ as half of the smallest input \textbf{used} count. After the radix point is chosen the inputs are split into lower and upper halves. Step 4 and 5 compute the lower halves. Step 6 and 7 computer the upper halves. After the halves have been computed the three intermediate half-size products must be computed. Step 8 and 9 compute the trivial products $x0 \cdot y0$ and $x1 \cdot y1$. The mp\_int $x0$ is used as a temporary variable after $x1 - x0$ has been computed. By using $x0$ instead of an additional temporary variable the algorithm can avoid an addition memory allocation operation. The remaining steps 13 through 18 compute the Karatsuba polynomial through a variety of digit shifting and addition operations. \index{bn\_mp\_karatsuba\_mul.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_mp\_karatsuba\_mul.c \vspace{-3mm} \begin{alltt} 016 017 /* c = |a| * |b| using Karatsuba Multiplication using 018 * three half size multiplications 019 * 020 * Let B represent the radix [e.g. 2**DIGIT_BIT] and 021 * let n represent half of the number of digits in 022 * the min(a,b) 023 * 024 * a = a1 * B**n + a0 025 * b = b1 * B**n + b0 026 * 027 * Then, a * b => 028 a1b1 * B**2n + ((a1 - a0)(b1 - b0) + a0b0 + a1b1) * B + a0b0 029 * 030 * Note that a1b1 and a0b0 are used twice and only need to be 031 * computed once. So in total three half size (half # of 032 * digit) multiplications are performed, a0b0, a1b1 and 033 * (a1-b1)(a0-b0) 034 * 035 * Note that a multiplication of half the digits requires 036 * 1/4th the number of single precision multiplications so in 037 * total after one call 25% of the single precision multiplications 038 * are saved. Note also that the call to mp_mul can end up back 039 * in this function if the a0, a1, b0, or b1 are above the threshold. 040 * This is known as divide-and-conquer and leads to the famous 041 * O(N**lg(3)) or O(N**1.584) work which is asymptopically lower than 042 * the standard O(N**2) that the baseline/comba methods use. 043 * Generally though the overhead of this method doesn't pay off 044 * until a certain size (N ~ 80) is reached. 045 */ 046 int 047 mp_karatsuba_mul (mp_int * a, mp_int * b, mp_int * c) 048 \{ 049 mp_int x0, x1, y0, y1, t1, x0y0, x1y1; 050 int B, err; 051 052 err = MP_MEM; 053 054 /* min # of digits */ 055 B = MIN (a->used, b->used); 056 057 /* now divide in two */ 058 B = B / 2; 059 060 /* init copy all the temps */ 061 if (mp_init_size (&x0, B) != MP_OKAY) 062 goto ERR; 063 if (mp_init_size (&x1, a->used - B) != MP_OKAY) 064 goto X0; 065 if (mp_init_size (&y0, B) != MP_OKAY) 066 goto X1; 067 if (mp_init_size (&y1, b->used - B) != MP_OKAY) 068 goto Y0; 069 070 /* init temps */ 071 if (mp_init_size (&t1, B * 2) != MP_OKAY) 072 goto Y1; 073 if (mp_init_size (&x0y0, B * 2) != MP_OKAY) 074 goto T1; 075 if (mp_init_size (&x1y1, B * 2) != MP_OKAY) 076 goto X0Y0; 077 078 /* now shift the digits */ 079 x0.sign = x1.sign = a->sign; 080 y0.sign = y1.sign = b->sign; 081 082 x0.used = y0.used = B; 083 x1.used = a->used - B; 084 y1.used = b->used - B; 085 086 \{ 087 register int x; 088 register mp_digit *tmpa, *tmpb, *tmpx, *tmpy; 089 090 /* we copy the digits directly instead of using higher level functions 091 * since we also need to shift the digits 092 */ 093 tmpa = a->dp; 094 tmpb = b->dp; 095 096 tmpx = x0.dp; 097 tmpy = y0.dp; 098 for (x = 0; x < B; x++) \{ 099 *tmpx++ = *tmpa++; 100 *tmpy++ = *tmpb++; 101 \} 102 103 tmpx = x1.dp; 104 for (x = B; x < a->used; x++) \{ 105 *tmpx++ = *tmpa++; 106 \} 107 108 tmpy = y1.dp; 109 for (x = B; x < b->used; x++) \{ 110 *tmpy++ = *tmpb++; 111 \} 112 \} 113 114 /* only need to clamp the lower words since by definition the 115 * upper words x1/y1 must have a known number of digits 116 */ 117 mp_clamp (&x0); 118 mp_clamp (&y0); 119 120 /* now calc the products x0y0 and x1y1 */ 121 /* after this x0 is no longer required, free temp [x0==t2]! */ 122 if (mp_mul (&x0, &y0, &x0y0) != MP_OKAY) 123 goto X1Y1; /* x0y0 = x0*y0 */ 124 if (mp_mul (&x1, &y1, &x1y1) != MP_OKAY) 125 goto X1Y1; /* x1y1 = x1*y1 */ 126 127 /* now calc x1-x0 and y1-y0 */ 128 if (mp_sub (&x1, &x0, &t1) != MP_OKAY) 129 goto X1Y1; /* t1 = x1 - x0 */ 130 if (mp_sub (&y1, &y0, &x0) != MP_OKAY) 131 goto X1Y1; /* t2 = y1 - y0 */ 132 if (mp_mul (&t1, &x0, &t1) != MP_OKAY) 133 goto X1Y1; /* t1 = (x1 - x0) * (y1 - y0) */ 134 135 /* add x0y0 */ 136 if (mp_add (&x0y0, &x1y1, &x0) != MP_OKAY) 137 goto X1Y1; /* t2 = x0y0 + x1y1 */ 138 if (mp_sub (&x0, &t1, &t1) != MP_OKAY) 139 goto X1Y1; /* t1 = x0y0 + x1y1 - (x1-x0)*(y1-y0) */ 140 141 /* shift by B */ 142 if (mp_lshd (&t1, B) != MP_OKAY) 143 goto X1Y1; /* t1 = (x0y0 + x1y1 - (x1-x0)*(y1-y0))<sign == b->sign) ? MP_ZPOS : MP_NEG; 023 024 if (MIN (a->used, b->used) >= TOOM_MUL_CUTOFF) \{ 025 res = mp_toom_mul(a, b, c); 026 \} else if (MIN (a->used, b->used) >= KARATSUBA_MUL_CUTOFF) \{ 027 res = mp_karatsuba_mul (a, b, c); 028 \} else \{ 029 030 /* can we use the fast multiplier? 031 * 032 * The fast multiplier can be used if the output will 033 * have less than MP_WARRAY digits and the number of 034 * digits won't affect carry propagation 035 */ 036 int digs = a->used + b->used + 1; 037 038 if ((digs < MP_WARRAY) && 039 MIN(a->used, b->used) <= 040 (1 << ((CHAR_BIT * sizeof (mp_word)) - (2 * DIGIT_BIT)))) \{ 041 res = fast_s_mp_mul_digs (a, b, c, digs); 042 \} else \{ 043 res = s_mp_mul (a, b, c); 044 \} 045 046 \} 047 c->sign = neg; 048 return res; 049 \} \end{alltt} \end{small} The implementation is rather simplistic and is not particularly noteworthy. Line 22 computes the sign of the result using the ``?'' operator from the C programming language. Line 40 computes $\delta$ using the fact that $1 << k$ is equal to $2^k$. \section{Squaring} Squaring is a special case of multiplication where both multiplicands are equal. At first it may seem like there is no significant optimization available but in fact there is. Consider the multiplication of $576$ against $241$. In total there will be nine single precision multiplications performed which are $1\cdot 6$, $1 \cdot 7$, $1 \cdot 5$, $4 \cdot 6$, $4 \cdot 7$, $4 \cdot 5$, $2 \cdot 6$, $2 \cdot 7$ and $2 \cdot 5$. Now consider the multiplication of $123$ against $123$. The nine products are $3 \cdot 3$, $3 \cdot 2$, $3 \cdot 1$, $2 \cdot 3$, $2 \cdot 2$, $2 \cdot 1$, $1 \cdot 3$, $1 \cdot 2$ and $1 \cdot 1$. On closer inspection some of the products are equivalent. For example, $3 \cdot 2 = 2 \cdot 3$ and $3 \cdot 1 = 1 \cdot 3$. For any $n$-digit input there are ${{\left (n^2 + n \right)}\over 2}$ possible unique single precision multiplications required. The following diagram demonstrates the operations required. \begin{figure}[here] \begin{center} \begin{tabular}{ccccc|c} &&1&2&3&\\ $\times$ &&1&2&3&\\ \hline && $3 \cdot 1$ & $3 \cdot 2$ & $3 \cdot 3$ & Row 0\\ & $2 \cdot 1$ & $2 \cdot 2$ & $2 \cdot 3$ && Row 1 \\ $1 \cdot 1$ & $1 \cdot 2$ & $1 \cdot 3$ &&& Row 2 \\ \end{tabular} \end{center} \caption{Squaring Optimization Diagram} \end{figure} Starting from zero and numbering the columns from right to left a very simple pattern becomes obvious. For the purposes of this discussion let $x$ represent the number being squared. The first observation is that in row $k$ the $2k$'th column of the product has a $\left (x_k \right)^2$ term in it. The second observation is that every column $j$ in row $k$ where $j \ne 2k$ is part of a double product. Every odd column is made up entirely of double products. In fact every column is made up of double products and at most one square (\textit{see the exercise section}). The third and final observation is that for row $k$ the first unique non-square term occurs at column $2k + 1$. For example, on row $1$ of the previous squaring, column one is part of the double product with column one from row zero. Column two of row one is a square and column three is the first unique column. \subsection{The Baseline Squaring Algorithm} The baseline squaring algorithm is meant to be a catch-all squaring algorithm. It will handle any of the input sizes that the faster routines will not handle. \begin{figure}[!here] \begin{small} \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{s\_mp\_sqr}. \\ \textbf{Input}. mp\_int $a$ \\ \textbf{Output}. $b \leftarrow a^2$ \\ \hline \\ 1. Init a temporary mp\_int of at least $2 \cdot a.used +1$ digits. (\textit{mp\_init\_size}) \\ 2. If step 1 failed return(\textit{MP\_MEM}) \\ 3. $t.used \leftarrow 2 \cdot a.used + 1$ \\ 4. For $ix$ from 0 to $a.used - 1$ do \\ \hspace{3mm}Calculate the square. \\ \hspace{3mm}4.1 $\hat r \leftarrow t_{2ix} + \left (a_{ix} \right )^2$ \\ \hspace{3mm}4.2 $t_{2ix} \leftarrow \hat r \mbox{ (mod }\beta\mbox{)}$ \\ \hspace{3mm}Calculate the double products after the square. \\ \hspace{3mm}4.3 $u \leftarrow \lfloor \hat r / \beta \rfloor$ \\ \hspace{3mm}4.4 For $iy$ from $ix + 1$ to $a.used - 1$ do \\ \hspace{6mm}4.4.1 $\hat r \leftarrow 2 \cdot a_{ix}a_{iy} + t_{ix + iy} + u$ \\ \hspace{6mm}4.4.2 $t_{ix + iy} \leftarrow \hat r \mbox{ (mod }\beta\mbox{)}$ \\ \hspace{6mm}4.4.3 $u \leftarrow \lfloor \hat r / \beta \rfloor$ \\ \hspace{3mm}Set the last carry. \\ \hspace{3mm}4.5 While $u > 0$ do \\ \hspace{6mm}4.5.1 $iy \leftarrow iy + 1$ \\ \hspace{6mm}4.5.2 $\hat r \leftarrow t_{ix + iy} + u$ \\ \hspace{6mm}4.5.3 $t_{ix + iy} \leftarrow \hat r \mbox{ (mod }\beta\mbox{)}$ \\ \hspace{6mm}4.5.4 $u \leftarrow \lfloor \hat r / \beta \rfloor$ \\ 5. Clamp excess digits of $t$. (\textit{mp\_clamp}) \\ 6. Exchange $b$ and $t$. \\ 7. Clear $t$ (\textit{mp\_clear}) \\ 8. Return(\textit{MP\_OKAY}) \\ \hline \end{tabular} \end{center} \end{small} \caption{Algorithm s\_mp\_sqr} \end{figure} \textbf{Algorithm s\_mp\_sqr.} This algorithm computes the square of an input using the three observations on squaring. It is based fairly faithfully on algorithm 14.16 of \cite[pp.596-597]{HAC}. Similar to algorithm s\_mp\_mul\_digs a temporary mp\_int is allocated to hold the result of the squaring. This allows the destination mp\_int to be the same as the source mp\_int without losing information part way through the squaring. The outer loop of this algorithm begins on step 4. It is best to think of the outer loop as walking down the rows of the partial results while the inner loop computes the columns of the partial result. Step 4.1 and 4.2 compute the square term for each row while step 4.3 and 4.4 propagate the carry and compute the double products. The requirement that a mp\_word be able to represent the range $0 \le x < 2 \beta^2$ arises from this very algorithm. The product $a_{ix}a_{iy}$ will lie in the range $0 \le x \le \beta^2 - 2\beta + 1$ which is obviously less than $\beta^2$ meaning that when it is multiply by two it can be represented by a mp\_word properly. Similar to algorithm s\_mp\_mul\_digs after every pass of the inner loop the destination is correctly set to the sum of all of the partial results calculated so far. This involves expensive carry propagation which will be eliminated shortly. \index{bn\_s\_mp\_sqr.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_s\_mp\_sqr.c \vspace{-3mm} \begin{alltt} 016 017 /* low level squaring, b = a*a, HAC pp.596-597, Algorithm 14.16 */ 018 int 019 s_mp_sqr (mp_int * a, mp_int * b) 020 \{ 021 mp_int t; 022 int res, ix, iy, pa; 023 mp_word r; 024 mp_digit u, tmpx, *tmpt; 025 026 pa = a->used; 027 if ((res = mp_init_size (&t, pa + pa + 1)) != MP_OKAY) \{ 028 return res; 029 \} 030 t.used = pa + pa + 1; 031 032 for (ix = 0; ix < pa; ix++) \{ 033 /* first calculate the digit at 2*ix */ 034 /* calculate double precision result */ 035 r = ((mp_word) t.dp[ix + ix]) + 036 ((mp_word) a->dp[ix]) * ((mp_word) a->dp[ix]); 037 038 /* store lower part in result */ 039 t.dp[ix + ix] = (mp_digit) (r & ((mp_word) MP_MASK)); 040 041 /* get the carry */ 042 u = (r >> ((mp_word) DIGIT_BIT)); 043 044 /* left hand side of A[ix] * A[iy] */ 045 tmpx = a->dp[ix]; 046 047 /* alias for where to store the results */ 048 tmpt = t.dp + (ix + ix + 1); 049 050 for (iy = ix + 1; iy < pa; iy++) \{ 051 /* first calculate the product */ 052 r = ((mp_word) tmpx) * ((mp_word) a->dp[iy]); 053 054 /* now calculate the double precision result, note we use 055 * addition instead of *2 since its easier to optimize 056 */ 057 r = ((mp_word) * tmpt) + r + r + ((mp_word) u); 058 059 /* store lower part */ 060 *tmpt++ = (mp_digit) (r & ((mp_word) MP_MASK)); 061 062 /* get carry */ 063 u = (r >> ((mp_word) DIGIT_BIT)); 064 \} 065 /* propagate upwards */ 066 while (u != ((mp_digit) 0)) \{ 067 r = ((mp_word) * tmpt) + ((mp_word) u); 068 *tmpt++ = (mp_digit) (r & ((mp_word) MP_MASK)); 069 u = (r >> ((mp_word) DIGIT_BIT)); 070 \} 071 \} 072 073 mp_clamp (&t); 074 mp_exch (&t, b); 075 mp_clear (&t); 076 return MP_OKAY; 077 \} \end{alltt} \end{small} Inside the outer loop (\textit{see line 32}) the square term is calculated on line 35. Line 42 extracts the carry from the square term. Aliases for $a_{ix}$ and $t_{ix+iy}$ are initialized on lines 45 and 48 respectively. The doubling is performed using two additions (\textit{see line 57}) since it is usually faster than shifting if not at least as fast. \subsection{Faster Squaring by the ``Comba'' Method} A major drawback to the baseline method is the requirement for single precision shifting inside the $O(n^2)$ work level. Squaring has an additional drawback that it must double the product inside the inner loop as well. As for multiplication the Comba technique can be used to eliminate these performance hazards. The first obvious solution is to make an array of mp\_words which will hold all of the columns. This will indeed eliminate all of the carry propagation operations from the inner loop. However, the inner product must still be doubled $O(n^2)$ times. The solution stems from the simple fact that $2a + 2b + 2c = 2(a + b + c)$. That is the sum of all of the double products is equal to double the sum of all the products. For example, $ab + ba + ac + ca = 2ab + 2ac = 2(ab + ac)$. However, we cannot simply double all of the columns since the squares appear only once per row. The most practical solution is to have two mp\_word arrays. One array will hold the squares and the other array will hold the double products. With both arrays the doubling and carry propagation can be moved to a $O(n)$ work level outside the $O(n^2)$ level. \newpage\begin{figure}[!here] \begin{small} \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{fast\_s\_mp\_sqr}. \\ \textbf{Input}. mp\_int $a$ \\ \textbf{Output}. $b \leftarrow a^2$ \\ \hline \\ Place two arrays of \textbf{MP\_WARRAY} mp\_words named $\hat W$ and $\hat {X}$ on the stack. \\ 1. If $b.alloc < 2a.used + 1$ then grow $b$ to $2a.used + 1$ digits. (\textit{mp\_grow}). \\ 2. If step 1 failed return(\textit{MP\_MEM}). \\ 3. for $ix$ from $0$ to $2a.used + 1$ do \\ \hspace{3mm}3.1 $\hat W_{ix} \leftarrow 0$ \\ \hspace{3mm}3.2 $\hat {X}_{ix} \leftarrow 0$ \\ 4. for $ix$ from $0$ to $a.used - 1$ do \\ \hspace{3mm}Compute the square.\\ \hspace{3mm}4.1 $\hat {X}_{ix+ix} \leftarrow \left ( a_ix \right )^2$ \\ \hspace{3mm}Compute the double products.\\ \hspace{3mm}4.2 for $iy$ from $ix + 1$ to $a.used - 1$ do \\ \hspace{6mm}4.2.1 $\hat W_{ix+iy} \leftarrow \hat W_{ix+iy} + a_{ix}a_{iy}$ \\ 5. $oldused \leftarrow b.used$ \\ 6. $b.used \leftarrow 2a.used + 1$ \\ Double the products and propagate the carries simultaneously. \\ 7. $\hat W_0 \leftarrow 2 \hat W_0 + \hat {X}_0$ \\ 8. for $ix$ from $1$ to $2a.used$ do \\ \hspace{3mm}8.1 $\hat W_{ix} \leftarrow 2 \hat W_{ix} + \hat {X}_{ix}$ \\ \hspace{3mm}8.2 $\hat W_{ix} \leftarrow \hat W_{ix} + \lfloor \hat W_{ix - 1} / \beta \rfloor$ \\ \hspace{3mm}8.3 $b_{ix-1} \leftarrow W_{ix-1} \mbox{ (mod }\beta\mbox{)}$ \\ 9. $b_{2a.used} \leftarrow \hat W_{2a.used} \mbox{ (mod }\beta\mbox{)}$ \\ 10. if $2a.used + 1 < oldused$ then do \\ \hspace{3mm}10.1 for $ix$ from $2a.used + 1$ to $oldused$ do \\ \hspace{6mm}10.1.1 $b_{ix} \leftarrow 0$ \\ 11. Clamp excess digits from $b$. (\textit{mp\_clamp}) \\ 12. Return(\textit{MP\_OKAY}). \\ \hline \end{tabular} \end{center} \end{small} \caption{Algorithm fast\_s\_mp\_sqr} \end{figure} \textbf{Algorithm fast\_s\_mp\_sqr.} This algorithm computes the square of an input using the Comba technique. It is designed to be a replacement for algorithm s\_mp\_sqr when the amount of input digits is less than \textbf{MP\_WARRAY} and less than $\delta \over 2$. This routine requires two arrays of mp\_words to be placed on the stack. The first array $\hat W$ will hold the double products and the second array $\hat X$ will hold the squares. Though only at most $MP\_WARRAY \over 2$ words of $\hat X$ are used it has proven faster on most processors to simply make it a full size array. The loop on step 3 will zero the two arrays to prepare them for the squaring step. Step 4.1 computes the squares of the product. Note how it simply assigns the value into the $\hat X$ array. The nested loop on step 4.2 computes the doubles of the products. In actuality that loop computes the sum of the products for each column. They are not doubled until later. After the squaring loop the products stored in $\hat W$ musted be doubled and the carries propagated forwards. It makes sense to do both operations at the same time. The expression $\hat W_{ix} \leftarrow 2 \hat W_{ix} + \hat {X}_{ix}$ computes the sum of the double product and the squares in place. \index{bn\_fast\_s\_mp\_sqr.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_fast\_s\_mp\_sqr.c \vspace{-3mm} \begin{alltt} 016 017 /* fast squaring 018 * 019 * This is the comba method where the columns of the product 020 * are computed first then the carries are computed. This 021 * has the effect of making a very simple inner loop that 022 * is executed the most 023 * 024 * W2 represents the outer products and W the inner. 025 * 026 * A further optimizations is made because the inner 027 * products are of the form "A * B * 2". The *2 part does 028 * not need to be computed until the end which is good 029 * because 64-bit shifts are slow! 030 * 031 * Based on Algorithm 14.16 on pp.597 of HAC. 032 * 033 */ 034 int 035 fast_s_mp_sqr (mp_int * a, mp_int * b) 036 \{ 037 int olduse, newused, res, ix, pa; 038 mp_word W2[MP_WARRAY], W[MP_WARRAY]; 039 040 /* calculate size of product and allocate as required */ 041 pa = a->used; 042 newused = pa + pa + 1; 043 if (b->alloc < newused) \{ 044 if ((res = mp_grow (b, newused)) != MP_OKAY) \{ 045 return res; 046 \} 047 \} 048 049 /* zero temp buffer (columns) 050 * Note that there are two buffers. Since squaring requires 051 * a outter and inner product and the inner product requires 052 * computing a product and doubling it (a relatively expensive 053 * op to perform n**2 times if you don't have to) the inner and 054 * outer products are computed in different buffers. This way 055 * the inner product can be doubled using n doublings instead of 056 * n**2 057 */ 058 memset (W, 0, newused * sizeof (mp_word)); 059 memset (W2, 0, newused * sizeof (mp_word)); 060 061 /* This computes the inner product. To simplify the inner N**2 loop 062 * the multiplication by two is done afterwards in the N loop. 063 */ 064 for (ix = 0; ix < pa; ix++) \{ 065 /* compute the outer product 066 * 067 * Note that every outer product is computed 068 * for a particular column only once which means that 069 * there is no need todo a double precision addition 070 */ 071 W2[ix + ix] = ((mp_word) a->dp[ix]) * ((mp_word) a->dp[ix]); 072 073 \{ 074 register mp_digit tmpx, *tmpy; 075 register mp_word *_W; 076 register int iy; 077 078 /* copy of left side */ 079 tmpx = a->dp[ix]; 080 081 /* alias for right side */ 082 tmpy = a->dp + (ix + 1); 083 084 /* the column to store the result in */ 085 _W = W + (ix + ix + 1); 086 087 /* inner products */ 088 for (iy = ix + 1; iy < pa; iy++) \{ 089 *_W++ += ((mp_word) tmpx) * ((mp_word) * tmpy++); 090 \} 091 \} 092 \} 093 094 /* setup dest */ 095 olduse = b->used; 096 b->used = newused; 097 098 /* now compute digits */ 099 \{ 100 register mp_digit *tmpb; 101 102 /* double first value, since the inner products are 103 * half of what they should be 104 */ 105 W[0] += W[0] + W2[0]; 106 107 tmpb = b->dp; 108 for (ix = 1; ix < newused; ix++) \{ 109 /* double/add next digit */ 110 W[ix] += W[ix] + W2[ix]; 111 112 W[ix] = W[ix] + (W[ix - 1] >> ((mp_word) DIGIT_BIT)); 113 *tmpb++ = (mp_digit) (W[ix - 1] & ((mp_word) MP_MASK)); 114 \} 115 /* set the last value. Note even if the carry is zero 116 * this is required since the next step will not zero 117 * it if b originally had a value at b->dp[2*a.used] 118 */ 119 *tmpb++ = (mp_digit) (W[(newused) - 1] & ((mp_word) MP_MASK)); 120 121 /* clear high digits */ 122 for (; ix < olduse; ix++) \{ 123 *tmpb++ = 0; 124 \} 125 \} 126 127 mp_clamp (b); 128 return MP_OKAY; 129 \} \end{alltt} \end{small} -- Write something deep and insightful later, Tom. \subsection{Polynomial Basis Squaring} The same algorithm that performs optimal polynomial basis multiplication can be used to perform polynomial basis squaring. The minor exception is that $\zeta_y = f(y) \cdot g(y)$ is actually equivalent to $\zeta_y = f(y)^2$ since $f(y) = g(y)$. That is instead of performing $2n + 1$ multiplications to find the $\zeta$ relations squaring operations are performed instead. \subsection{Karatsuba Squaring} Let $f(x) = ax + b$ represent the polynomial basis representation of a number to square. Let $h(x) = \left ( f(x) \right )^2$ represent the square of the polynomial. The Karatsuba equation can be modified to square a number with the following equation. \begin{equation} h(x) = a^2x^2 + \left (a^2 + b^2 - (a - b)^2 \right )x + b^2 \end{equation} Upon closer inspection this equation only requires the calculation of three half-sized squares: $a^2$, $b^2$ and $(a - b)^2$. As in Karatsuba multiplication this algorithm can be applied recursively on the input and will achieve an asymptotic running time of $O \left ( n^{lg(3)} \right )$. If the asymptotic time of Karatsuba squaring and multiplication is the same why not simply use the multiplication algorithm instead? The answer to this question arises from the cutoff point for squaring. As in multiplication there exists a cutoff point at which the time required for a Comba based squaring and a Karatsuba based squaring meet. Due to the overhead inherent in the Karatsuba method the cutoff point is fairly high. For example, on an Athlon processor with $\beta = 2^{28}$ the cutoff point is around 127 digits. Consider squaring a 200 digit number with this technique. It will be split into two 100 digit halves which are subsequently squared. The 100 digit numbers will not be squared using Karatsuba but instead the faster Comba based squaring algorithm. If Karatsuba multiplication were used instead the 100 digit numbers would be squared with a slower Comba based multiplication. \newpage\begin{figure}[!here] \begin{small} \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{mp\_karatsuba\_sqr}. \\ \textbf{Input}. mp\_int $a$ \\ \textbf{Output}. $b \leftarrow a^2$ \\ \hline \\ 1. Initialize the following temporary mp\_ints: $x0$, $x1$, $t1$, $t2$, $x0x0$ and $x1x1$. \\ 2. If any of the initializations on step 1 failed return(\textit{MP\_MEM}). \\ \\ Split the input. e.g. $a = x1\beta^B + x0$ \\ 3. $B \leftarrow a.used / 2$ \\ 4. $x0 \leftarrow a \mbox{ (mod }\beta^B\mbox{)}$ (\textit{mp\_mod\_2d}) \\ 5. $x1 \leftarrow \lfloor a / \beta^B \rfloor$ (\textit{mp\_lshd}) \\ \\ Calculate the three squares. \\ 6. $x0x0 \leftarrow x0^2$ (\textit{mp\_sqr}) \\ 7. $x1x1 \leftarrow x1^2$ \\ 8. $t1 \leftarrow x1 - x0$ (\textit{mp\_sub}) \\ 9. $t1 \leftarrow t1^2$ \\ \\ Compute the middle term. \\ 10. $t2 \leftarrow x0x0 + x1x1$ (\textit{s\_mp\_add}) \\ 11. $t1 \leftarrow t2 - t1$ \\ \\ Compute final product. \\ 12. $t1 \leftarrow t1\beta^B$ (\textit{mp\_lshd}) \\ 13. $x1x1 \leftarrow x1x1\beta^{2B}$ \\ 14. $t1 \leftarrow t1 + x0x0$ \\ 15. $b \leftarrow t1 + x1x1$ \\ 16. Return(\textit{MP\_OKAY}). \\ \hline \end{tabular} \end{center} \end{small} \caption{Algorithm mp\_karatsuba\_sqr} \end{figure} \textbf{Algorithm mp\_karatsuba\_sqr.} This algorithm computes the square of an input $a$ using the Karatsuba technique. This algorithm is very much similar to the Karatsuba based multiplication algorithm. The radix point for squaring is simply the placed above the median of the digits. Step 3, 4 and 5 compute the two halves required using $B$ as the radix point. The first two squares in steps 6 and 7 are rather straightforward while the last square is in a more compact form. By expanding $\left (x1 - x0 \right )^2$ the $x1^2$ and $x0^2$ terms in the middle disappear, that is $x1^2 + x0^2 - (x1 - x0)^2 = 2 \cdot x0 \cdot x1$. Now if $5n$ single precision additions and a squaring of $n$-digits is faster than multiplying two $n$-digit numbers and doubling then this method is faster. Assuming no further recursions occur the difference can be estimated. Let $p$ represent the cost of a single precision addition and $q$ the cost of a single precision multiplication both in terms of time\footnote{Or machine clock cycles.}. The question reduces to whether the following equation is true or not. \begin{equation} 5np +{{q(n^2 + n)} \over 2} \le pn + qn^2 \end{equation} For example, on an AMD Athlon processor $p = {1 \over 3}$ and $q = 6$. This implies that the following inequality should hold. \begin{center} \begin{tabular}{rcl} $5n + 3n^2 + 3n$ & $<$ & ${n \over 3} + 6n^2$ \\ ${25 \over 3} + 3n$ & $<$ & ${1 \over 3} + 6n$ \\ ${25 \over 3}$ & $<$ & $3n$ \\ ${25 \over 9}$ & $<$ & $n$ \\ \end{tabular} \end{center} This results in a cutoff point around $n = 3$. As a consequence it is actually faster to compute the middle term the ``long way'' on processors where multiplication is substantially slower\footnote{On the Athlon there is a 1:17 ratio between clock cycles for addition and multiplication. On the Intel P4 processor this ratio is 1:29 making this method even more beneficial. The only common exception is the ARMv4 processor which has a ratio of 1:7. } than simpler operations such as addition. \index{bn\_mp\_karatsuba\_sqr.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_mp\_karatsuba\_sqr.c \vspace{-3mm} \begin{alltt} 016 017 /* Karatsuba squaring, computes b = a*a using three 018 * half size squarings 019 * 020 * See comments of mp_karatsuba_mul for details. It 021 * is essentially the same algorithm but merely 022 * tuned to perform recursive squarings. 023 */ 024 int 025 mp_karatsuba_sqr (mp_int * a, mp_int * b) 026 \{ 027 mp_int x0, x1, t1, t2, x0x0, x1x1; 028 int B, err; 029 030 err = MP_MEM; 031 032 /* min # of digits */ 033 B = a->used; 034 035 /* now divide in two */ 036 B = B / 2; 037 038 /* init copy all the temps */ 039 if (mp_init_size (&x0, B) != MP_OKAY) 040 goto ERR; 041 if (mp_init_size (&x1, a->used - B) != MP_OKAY) 042 goto X0; 043 044 /* init temps */ 045 if (mp_init_size (&t1, a->used * 2) != MP_OKAY) 046 goto X1; 047 if (mp_init_size (&t2, a->used * 2) != MP_OKAY) 048 goto T1; 049 if (mp_init_size (&x0x0, B * 2) != MP_OKAY) 050 goto T2; 051 if (mp_init_size (&x1x1, (a->used - B) * 2) != MP_OKAY) 052 goto X0X0; 053 054 \{ 055 register int x; 056 register mp_digit *dst, *src; 057 058 src = a->dp; 059 060 /* now shift the digits */ 061 dst = x0.dp; 062 for (x = 0; x < B; x++) \{ 063 *dst++ = *src++; 064 \} 065 066 dst = x1.dp; 067 for (x = B; x < a->used; x++) \{ 068 *dst++ = *src++; 069 \} 070 \} 071 072 x0.used = B; 073 x1.used = a->used - B; 074 075 mp_clamp (&x0); 076 077 /* now calc the products x0*x0 and x1*x1 */ 078 if (mp_sqr (&x0, &x0x0) != MP_OKAY) 079 goto X1X1; /* x0x0 = x0*x0 */ 080 if (mp_sqr (&x1, &x1x1) != MP_OKAY) 081 goto X1X1; /* x1x1 = x1*x1 */ 082 083 /* now calc (x1-x0)**2 */ 084 if (mp_sub (&x1, &x0, &t1) != MP_OKAY) 085 goto X1X1; /* t1 = x1 - x0 */ 086 if (mp_sqr (&t1, &t1) != MP_OKAY) 087 goto X1X1; /* t1 = (x1 - x0) * (x1 - x0) */ 088 089 /* add x0y0 */ 090 if (s_mp_add (&x0x0, &x1x1, &t2) != MP_OKAY) 091 goto X1X1; /* t2 = x0x0 + x1x1 */ 092 if (mp_sub (&t2, &t1, &t1) != MP_OKAY) 093 goto X1X1; /* t1 = x0x0 + x1x1 - (x1-x0)*(x1-x0) */ 094 095 /* shift by B */ 096 if (mp_lshd (&t1, B) != MP_OKAY) 097 goto X1X1; /* t1 = (x0x0 + x1x1 - (x1-x0)*(x1-x0))<used >= TOOM_SQR_CUTOFF) \{ 023 res = mp_toom_sqr(a, b); 024 \} else if (a->used >= KARATSUBA_SQR_CUTOFF) \{ 025 res = mp_karatsuba_sqr (a, b); 026 \} else \{ 027 028 /* can we use the fast multiplier? */ 029 if ((a->used * 2 + 1) < MP_WARRAY && 030 a->used < 031 (1 << (sizeof(mp_word) * CHAR_BIT - 2*DIGIT_BIT - 1))) \{ 032 res = fast_s_mp_sqr (a, b); 033 \} else \{ 034 res = s_mp_sqr (a, b); 035 \} 036 \} 037 b->sign = MP_ZPOS; 038 return res; 039 \} \end{alltt} \end{small} \section*{Exercises} \begin{tabular}{cl} $\left [ 3 \right ] $ & Devise an efficient algorithm for selection of the radix point to handle inputs \\ & that have different number of digits in Karatsuba multiplication. \\ & \\ $\left [ 3 \right ] $ & In section 6.3 the fact that every column of a squaring is made up \\ & of double products and at most one square is stated. Prove this statement. \\ & \\ $\left [ 2 \right ] $ & In the Comba squaring algorithm half of the $\hat X$ variables are not used. \\ & Revise algorithm fast\_s\_mp\_sqr to shrink the $\hat X$ array. \\ & \\ $\left [ 3 \right ] $ & Prove the equation for Karatsuba squaring. \\ & \\ $\left [ 1 \right ] $ & Prove that Karatsuba squaring requires $O \left (n^{lg(3)} \right )$ time. \\ & \\ $\left [ 2 \right ] $ & Determine the minimal ratio between addition and multiplication clock cycles \\ & required for equation $6.7$ to be true. \\ & \\ \end{tabular} \chapter{Modular Reduction} \section{Basics of Modular Reduction} \index{modular residue} Modular reduction is an operation that arises quite often within public key cryptography algorithms. A number is said to be reduced modulo another number by finding the remainder of division. If an integer $a$ is reduced modulo $b$ that is to solve the equation $a = bq + p$ then $p$ is the result. To phrase that another way ``$p$ is congruent to $a$ modulo $b$'' which is also written as $p \equiv a \mbox{ (mod }b\mbox{)}$. In other vernacular $p$ is known as the ``modular residue'' which leads to ``quadratic residue''\footnote{That's fancy talk for $b \equiv a^2 \mbox{ (mod }p\mbox{)}$.} and other forms of residues. \index{modulus} Modular reductions are normally used to form finite groups such as fields and rings. For example, in the RSA public key algorithm \cite{RSAPAPER} two private primes $p$ and $q$ are chosen which when multiplied $n = pq$ forms a composite modulus. When operations such as multiplication and squaring are performed on units of the ring $\Z_n$ a finite multiplicative sub-group is formed. This sub-group is the group used to perform RSA operations. Do not worry to much about how RSA works as it is not important for this discussion. The most common usage for performance driven modular reductions is in modular exponentiation algorithms. That is to compute $d = a^b \mbox{ (mod }c\mbox{)}$ as fast as possible. As will be discussed in the subsequent chapter there exists fast algorithms for computing modular exponentiations without having to perform (\textit{in this example}) $b$ multiplications. These algorithms will produce partial results in the range $0 \le x < c^2$ which can be taken advantage of. The obvious line of thinking is to use an integer division routine and just extract the remainder. While this is equivalent to finding the modular residue it turns out that the limited range of the input can be exploited to create several efficient algorithms. \section{The Barrett Reduction} The Barrett reduction algorithm \cite{BARRETT} was inspired by fast division algorithms which multiply by the reciprocal to emulate division. Barretts observation was that the residue $c$ of $a$ modulo $b$ is equal to \begin{equation} c = a - b \cdot \lfloor a/b \rfloor \end{equation} Since algorithms such as modular reduction would be using the same modulus extensively, using typical DSP intuition the next step would be to replace $a/b$ with a multiplication by the reciprocal. However, DSP intuition on its own will not work as these numbers are considerably larger than the precision of common DSP floating point data types. It would take another common optimization to optimize the algorithm. \subsection{Fixed Point Arithmetic} The trick used to optimize the above equation is based on a technique of emulating floating point data types with fixed precision integers. Fixed point arithmetic would be vastly popularlized in the mid 1990s for bringing 3d-games to the mass market. The idea is to take a normal $k$-bit integer data type and break it into $p$-bit integer and a $q$-bit fraction part (\textit{where $p+q = k$}). In this system a $k$-bit integer $n$ would actually represent $n/2^q$. For example, with $q = 4$ the integer $n = 37$ would actually represent the value $2.3125$. To multiply two fixed point numbers the integers are multiplied using traditional arithmetic and subsequently normalized. For example, with $q = 4$ to multiply the integers $9$ and $5$ they must be converted to fixed point first by multiplying by $2^q$. Let $a = 9(2^q)$ represent the fixed point representation of $9$ and $b = 5(2^q)$ represent the fixed point representation of $5$. The product $ab$ is equal to $45(2^{2q})$ which when normalized produces $45(2^q)$. Using fixed point arithmetic division can be easily achieved by multiplying by the reciprocal. If $2^q$ is equivalent to one than $2^q/b$ is equivalent to $1/b$ using real arithmetic. Using this fact dividing an integer $a$ by another integer $b$ can be achieved with the following expression. \begin{equation} \lfloor (a \cdot (\lfloor 2^q / b \rfloor))/2^q \rfloor \end{equation} The precision of the division is proportional to the value of $q$. If the divisor $b$ is used frequently as is the case with modular exponentiation pre-computing $2^q/b$ will allow a division to be performed with a multiplication and a right shift. Both operations are considerably faster than division on most processors. Consider dividing $19$ by $5$. The correct result is $\lfloor 19/5 \rfloor = 3$. With $q = 3$ the reciprocal is $\lfloor 2^q/5 \rfloor = 1$ which leads to a product of $19$ which when divided by $2^q$ produces $2$. However, with $q = 4$ the reciprocal is $\lfloor 2^q/5 \rfloor = 3$ and the result of the emulated division is $\lfloor 3 \cdot 19 / 2^q \rfloor = 3$ which is correct. Plugging this form of divison into the original equation the following modular residue equation arises. \begin{equation} c = a - b \cdot \lfloor (a \cdot (\lfloor 2^q / b \rfloor))/2^q \rfloor \end{equation} Using the notation from \cite{BARRETT} the value of $\lfloor 2^q / b \rfloor$ will be represented by the $\mu$ symbol. Using the $\mu$ variable also helps re-inforce the idea that it is meant to be computed once and re-used. \begin{equation} c = a - b \cdot \lfloor (a \cdot \mu)/2^q \rfloor \end{equation} Provided that $2^q > b^2$ this algorithm will produce a quotient that is either exactly correct or off by a value of one. Let $n$ represent the number of digits in $b$. This algorithm requires approximately $2n^2$ single precision multiplications to produce the quotient and another $n^2$ single precision multiplications to find the residue. In total $3n^2$ single precision multiplications are required to reduce the number. For example, if $b = 1179677$ and $q = 41$ ($2^q > b^2$), then the reciprocal $\mu$ is equal to $\lfloor 2^q / b \rfloor = 1864089$. Consider reducing $a = 180388626447$ modulo $b$ using the above reduction equation. Using long division the quotient $\lfloor a/b \rfloor$ is equal to the quotient found using the fixed point method. In this case the quotient is $\lfloor (a \cdot \mu)/2^q \rfloor = 152913$ and can produce the modular residue $a - 152913b = 677346$. \subsection{Choosing a Radix Point} Using the fixed point representation a modular reduction can be performed with $3n^2$ single precision multiplications. If that were the best that could be achieved a full division might as well be used in its place. The key to optimizing the reduction is to reduce the precision of the initial multiplication that finds the quotient. Let $a$ represent the number of which the residue is sought. Let $b$ represent the modulus used to find the residue. Let $m$ represent the number of digits in $b$. For the purposes of this discussion we will assume that the number of digits in $a$ is $2m$. Dividing $a$ by $b$ is the same as dividing a $2m$ digit integer by a $m$ digit integer. Digits below the $m - 1$'th digit of $a$ will contribute at most a value of $1$ to the quotient because $\beta^k < b$ for any $0 \le k \le m - 1$. Since those digits do not contribute much to the quotient the observation is that they might as well be zero. However, if the digits ``might as well be zero'' they might as well not be there in the first place. Let $q_0 = \lfloor a/\beta^{m-1} \rfloor$ represent the input with the zeroes trimmed. Now the modular reduction is trimmed to the almost equivalent equation \begin{equation} c = a - b \cdot \lfloor (q_0 \cdot \mu) / \beta^{m+1} \rfloor \end{equation} Notice how the original divisor $2^q$ has been replaced with $\beta^{m+1}$. Also note how the exponent on the divisor $m+1$ when added to the amount $q_0$ was shifted by ($m-1$) equals $2m$. If the optimization had not been performed the divisor would have the exponent $2m$ so in the end the exponents do ``add up''. By using whole digits the algorithm is much faster since shifting digits is typically slower than simply copying them. Using the above equation the quotient $\lfloor (q_0 \cdot \mu) / \beta^{m+1} \rfloor$ can be off from the true quotient by at most two implying that $0 \le a - b \cdot \lfloor (q_0 \cdot \mu) / \beta^{m+1} \rfloor < 3b$. By first subtracting $b$ times the quotient and then conditionally subtracting $b$ once or twice the residue is found. The quotient is now found using $(m + 1)(m) = m^2 + m$ single precision multiplications and the residue with an additional $m^2$ single precision multiplications. In total $2m^2 + m$ single precision multiplications are required which is considerably faster than the original attempt. For example, let $\beta = 10$ represent the radix of the digits. Let $b = 9999$ represent the modulus which implies $m = 4$. Let $a = 99929878$ represent the value of which the residue is desired. In this case $q = 10$ which means that $\mu = \lfloor \beta^{2m}/b \rfloor = 10001$. With this optimization the multiplicand for the quotient is $q_0 = \lfloor a / \beta^{m - 1} \rfloor = 99929$. The quotient is then $\lfloor (q_0 \cdot \mu) / \beta^{m+1} \rfloor = 9993$. Subtracting $9993b$ from $a$ and the correct residue $9871 \equiv a \mbox{ (mod }b\mbox{)}$ is found. \subsection{Trimming the Quotient} So far the reduction algorithm has been optimized from $3m^2$ single precision multiplications down to $2m^2 + m$ single precision multiplications. As it stands now the algorithm is already fairly fast compared to a full integer division algorithm. However, there is still room for optimization. After the first multiplication inside the quotient ($q_0 \cdot \mu$) the value is shifted right by $m + 1$ places effectively nullifying the lower half of the product. It would be nice to be able to remove those digits from the product to effectively cut down the number of multiplications. If the number of digits in the modulus $m$ is far less than $\beta$ a full product is not required. In fact the lower $m - 2$ digits will not affect the upper half of the product at all and do not need to be computed. The value of $\mu$ is a $m$-digit number and $q_0$ is a $m + 1$ digit number. Using a full multiplier $(m + 1)(m) = m^2 + m$ single precision multiplications would be required. Using a multiplier that will only produce digits at and above the $m - 1$'th digit reduces the number of single precision multiplications to ${m^2 + m} \over 2$ single precision multiplications. \subsection{Trimming the Residue} After the quotient has been calculated it is used to reduce the input. As previously noted the algorithm is not exact and it can be off by a small multiple of the modulus, that is $0 \le a - b \cdot \lfloor (q_0 \cdot \mu) / \beta^{m+1} \rfloor < 3b$. If $b$ is $m$ digits than the result of reduction equation is a value of at most $m + 1$ digits (\textit{provided $3 < \beta$}) implying that the upper $m - 1$ digits are implicitly zero. The next optimization arises from this very fact. Instead of computing $b \cdot \lfloor (q_0 \cdot \mu) / \beta^{m+1} \rfloor$ using a full $O(m^2)$ multiplication algorithm only the lower $m+1$ digits of the product have to be computed. Similarly the value of $a$ can be reduced modulo $\beta^{m+1}$ before the multiple of $b$ is subtracted which simplifes the subtraction as well. A multiplication that produces only the lower $m+1$ digits requires ${m^2 + 3m - 2} \over 2$ single precision multiplications. With both optimizations in place the algorithm is the algorithm Barrett proposed. It requires $m^2 + 2m - 1$ single precision multiplications which is considerably faster than the straightforward $3m^2$ method. \subsection{The Barrett Algorithm} \newpage\begin{figure}[!here] \begin{small} \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{mp\_reduce}. \\ \textbf{Input}. mp\_int $a$, mp\_int $b$ and $\mu = \lfloor \beta^{2m}/b \rfloor$ $(0 \le a < b^2, b > 1)$ \\ \textbf{Output}. $c \leftarrow a \mbox{ (mod }b\mbox{)}$ \\ \hline \\ Let $m$ represent the number of digits in $b$. \\ 1. Make a copy of $a$ and store it in $q$. (\textit{mp\_init\_copy}) \\ 2. $q \leftarrow \lfloor q / \beta^{m - 1} \rfloor$ (\textit{mp\_rshd}) \\ \\ Produce the quotient. \\ 3. $q \leftarrow q \cdot \mu$ (\textit{note: only produce digits at or above $m-1$}) \\ 4. $q \leftarrow \lfloor q / \beta^{m + 1} \rfloor$ \\ \\ Subtract the multiple of modulus from the input. \\ 5. $c \leftarrow a \mbox{ (mod }\beta^{m+1}\mbox{)}$ (\textit{mp\_mod\_2d}) \\ 6. $q \leftarrow q \cdot b \mbox{ (mod }\beta^{m+1}\mbox{)}$ (\textit{s\_mp\_mul\_digs}) \\ 7. $c \leftarrow c - q$ (\textit{mp\_sub}) \\ \\ Add $\beta^{m+1}$ if a carry occured. \\ 8. If $c < 0$ then (\textit{mp\_cmp\_d}) \\ \hspace{3mm}8.1 $q \leftarrow 1$ (\textit{mp\_set}) \\ \hspace{3mm}8.2 $q \leftarrow q \cdot \beta^{m+1}$ (\textit{mp\_lshd}) \\ \hspace{3mm}8.3 $c \leftarrow c + q$ \\ \\ Now subtract the modulus if the residue is too large (e.g. quotient too small). \\ 9. While $c \ge b$ do (\textit{mp\_cmp}) \\ \hspace{3mm}9.1 $c \leftarrow c - b$ \\ 10. Clear $q$. \\ 11. Return(\textit{MP\_OKAY}) \\ \hline \end{tabular} \end{center} \end{small} \caption{Algorithm mp\_reduce} \end{figure} \textbf{Algorithm mp\_reduce.} This algorithm will reduce the input $a$ modulo $b$ in place using the Barrett algorithm. It is loosely based on algorithm 14.42 of \cite[pp. 602]{HAC} which is based on \cite{BARRETT}. The algorithm has several restrictions and assumptions which must be adhered to for the algorithm to work. First the modulus $b$ is assumed to be positive and greater than one. If the modulus were less than or equal to one than subtracting a multiple of it would either accomplish nothing or actually enlarge the input. The input $a$ must be in the range $0 \le a < b^2$ in order for the quotient to have enough precision. Technically the algorithm will still work if $a \ge b^2$ but it will take much longer to finish. The value of $\mu$ is passed as an argument to this algorithm and is assumed to be setup before the algorithm is used. Recall that the multiplication for the quotient on step 3 must only produce digits at or above the $m-1$'th position. An algorithm called $s\_mp\_mul\_high\_digs$ which has not been presented is used to accomplish this task. This optimal algorithm can only be used if the number of digits in $b$ is very much smaller than $\beta$. After the multiple of the modulus has been subtracted from $a$ the residue must be fixed up in case its negative. While it is known that $a \ge b \cdot \lfloor (q_0 \cdot \mu) / \beta^{m+1} \rfloor$ only the lower $m+1$ digits are being used to compute the residue. In this case the invariant $\beta^{m+1}$ must be added to the residue to make it positive again. The while loop at step 9 will subtract $b$ until the residue is less than $b$. If the algorithm is performed correctly this step is only performed upto two times. However, if $a \ge b^2$ than it will iterate substantially more times than it should. \index{bn\_mp\_reduce.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_mp\_reduce.c \vspace{-3mm} \begin{alltt} 016 017 /* reduces x mod m, assumes 0 < x < m**2, mu is 018 * precomputed via mp_reduce_setup. 019 * From HAC pp.604 Algorithm 14.42 020 */ 021 int 022 mp_reduce (mp_int * x, mp_int * m, mp_int * mu) 023 \{ 024 mp_int q; 025 int res, um = m->used; 026 027 /* q = x */ 028 if ((res = mp_init_copy (&q, x)) != MP_OKAY) \{ 029 return res; 030 \} 031 032 /* q1 = x / b**(k-1) */ 033 mp_rshd (&q, um - 1); 034 035 /* according to HAC this is optimization is ok */ 036 if (((unsigned long) m->used) > (((mp_digit)1) << (DIGIT_BIT - 1))) \{ 037 if ((res = mp_mul (&q, mu, &q)) != MP_OKAY) \{ 038 goto CLEANUP; 039 \} 040 \} else \{ 041 if ((res = s_mp_mul_high_digs (&q, mu, &q, um - 1)) != MP_OKAY) \{ 042 goto CLEANUP; 043 \} 044 \} 045 046 /* q3 = q2 / b**(k+1) */ 047 mp_rshd (&q, um + 1); 048 049 /* x = x mod b**(k+1), quick (no division) */ 050 if ((res = mp_mod_2d (x, DIGIT_BIT * (um + 1), x)) != MP_OKAY) \{ 051 goto CLEANUP; 052 \} 053 054 /* q = q * m mod b**(k+1), quick (no division) */ 055 if ((res = s_mp_mul_digs (&q, m, &q, um + 1)) != MP_OKAY) \{ 056 goto CLEANUP; 057 \} 058 059 /* x = x - q */ 060 if ((res = mp_sub (x, &q, x)) != MP_OKAY) \{ 061 goto CLEANUP; 062 \} 063 064 /* If x < 0, add b**(k+1) to it */ 065 if (mp_cmp_d (x, 0) == MP_LT) \{ 066 mp_set (&q, 1); 067 if ((res = mp_lshd (&q, um + 1)) != MP_OKAY) 068 goto CLEANUP; 069 if ((res = mp_add (x, &q, x)) != MP_OKAY) 070 goto CLEANUP; 071 \} 072 073 /* Back off if it's too big */ 074 while (mp_cmp (x, m) != MP_LT) \{ 075 if ((res = s_mp_sub (x, m, x)) != MP_OKAY) \{ 076 break; 077 \} 078 \} 079 080 CLEANUP: 081 mp_clear (&q); 082 083 return res; 084 \} \end{alltt} \end{small} The first multiplication that determines the quotient can be performed by only producing the digits from $m - 1$ and up. This essentially halves the number of single precision multiplications required. However, the optimization is only safe if $\beta$ is much larger than the number of digits in the modulus. In the source code this is evaluated on lines 36 to 44 where algorithm s\_mp\_mul\_high\_digs is used when it is safe to do so. \subsection{The Barrett Setup Algorithm} In order to use algorithm mp\_reduce the value of $\mu$ must be calculated in advance. Ideally this value should be computed once and stored for future use so that the Barrett algorithm can be used without delay. \begin{figure}[!here] \begin{small} \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{mp\_reduce\_setup}. \\ \textbf{Input}. mp\_int $a$ ($a > 1$) \\ \textbf{Output}. $\mu \leftarrow \lfloor \beta^{2m}/a \rfloor$ \\ \hline \\ 1. $\mu \leftarrow 2^{2 \cdot lg(\beta) \cdot m}$ (\textit{mp\_2expt}) \\ 2. $\mu \leftarrow \lfloor \mu / b \rfloor$ (\textit{mp\_div}) \\ 3. Return(\textit{MP\_OKAY}) \\ \hline \end{tabular} \end{center} \end{small} \caption{Algorithm mp\_reduce\_setup} \end{figure} \textbf{Algorithm mp\_reduce\_setup.} This algorithm computes the reciprocal $\mu$ required for Barrett reduction. First $\beta^{2m}$ is calculated as $2^{2 \cdot lg(\beta) \cdot m}$ which is equivalent and much faster. The final value is computed by taking the integer quotient of $\lfloor \mu / b \rfloor$. \index{bn\_mp\_reduce\_setup.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_mp\_reduce\_setup.c \vspace{-3mm} \begin{alltt} 016 017 /* pre-calculate the value required for Barrett reduction 018 * For a given modulus "b" it calulates the value required in "a" 019 */ 020 int 021 mp_reduce_setup (mp_int * a, mp_int * b) 022 \{ 023 int res; 024 025 if ((res = mp_2expt (a, b->used * 2 * DIGIT_BIT)) != MP_OKAY) \{ 026 return res; 027 \} 028 return mp_div (a, b, a, NULL); 029 \} \end{alltt} \end{small} This simple routine calculates the reciprocal $\mu$ required by Barrett reduction. Note the extended usage of algorithm mp\_div where the variable which would received the remainder is passed as NULL. As will be discussed in section 9.1 the division routine allows both the quotient and the remainder to be passed as NULL meaning to ignore the value. \section{The Montgomery Reduction} Montgomery reduction\footnote{Thanks to Niels Ferguson for his insightful explanation of the algorithm.} \cite{MONT} is by far the most interesting form of reduction in common use. It computes a modular residue which is not actually equal to the residue of the input yet instead equal to a residue times a constant. However, as perplexing as this may sound the algorithm is relatively simple and very efficient. Throughout this entire section the variable $n$ will represent the modulus used to form the residue. As will be discussed shortly the value of $n$ must be odd. The variable $x$ will represent the quantity of which the residue is sought. Similar to the Barrett algorithm the input is restricted to $0 \le x < n^2$. To begin the description some simple number theory facts must be established. \textbf{Fact 1.} Adding $n$ to $x$ does not change the residue since in effect it adds one to the quotient $\lfloor x / n \rfloor$. \textbf{Fact 2.} If $x$ is even then performing a division by two in $\Z$ is congruent to $x \cdot 2^{-1} \mbox{ (mod }n\mbox{)}$. For example, if $n = 7$ and $x = 6$ then $x/2 = 3$. Using the modular inverse of two the same result is found. That is, $2^{-1} \equiv (n+1)/2 \equiv 4$ and $4 \cdot 6 \equiv 3 \mbox{ (mod }n\mbox{)}$. From these two simple facts the following simple algorithm can be derived. \newpage\begin{figure}[!here] \begin{small} \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{Montgomery Reduction}. \\ \textbf{Input}. Integer $x$, $n$ and $k$ \\ \textbf{Output}. $2^{-k}x \mbox{ (mod }n\mbox{)}$ \\ \hline \\ 1. for $t$ from $1$ to $k$ do \\ \hspace{3mm}1.1 If $x$ is odd then \\ \hspace{6mm}1.1.1 $x \leftarrow x + n$ \\ \hspace{3mm}1.2 $x \leftarrow x/2$ \\ 2. Return $x$. \\ \hline \end{tabular} \end{center} \end{small} \caption{Algorithm Montgomery Reduction} \end{figure} The algorithm reduces the input one bit at a time using the two congruencies stated previously. Inside the loop $n$, which is odd, is added to $x$ if $x$ is odd. This forces $x$ to be even which allows the division by two in $\Z$ to be congruent to a modular division by two. Let $r$ represent the final result of the Montgomery algorithm. If $k > lg(n)$ and $0 \le x < n^2$ then the final result is limited to $0 \le r < \lfloor x/2^k \rfloor + n$. As a result at most a single subtraction is required to get the residue desired. Let $k = \lfloor lg(n) \rfloor + 1$ represent the number of bits in $n$. The current algorithm requires $2k^2$ single precision shifts and $k^2$ single precision additions. At this rate the algorithm is most certainly slower than Barrett reduction and not terribly useful. Fortunately there exists an alternative representation of the algorithm. \begin{figure}[!here] \begin{small} \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{Montgomery Reduction} (modified I). \\ \textbf{Input}. Integer $x$, $n$ and $k$ \\ \textbf{Output}. $2^{-k}x \mbox{ (mod }n\mbox{)}$ \\ \hline \\ 1. for $t$ from $0$ to $k - 1$ do \\ \hspace{3mm}1.1 If the $t$'th bit of $x$ is one then \\ \hspace{6mm}1.1.1 $x \leftarrow x + 2^tn$ \\ 2. Return $x/2^k$. \\ \hline \end{tabular} \end{center} \end{small} \caption{Algorithm Montgomery Reduction (modified I)} \end{figure} This algorithm is equivalent since $2^tn$ is a multiple of $n$ and the lower $k$ bits of $x$ are zero by step 2. The number of single precision shifts has now been reduced from $2k^2$ to $k^2 + 1$ which is only a small improvement. \subsection{Digit Based Montgomery Reduction} Instead of computing the reduction on a bit-by-bit basis it is actually much faster to compute it on digit-by-digit basis. Consider the previous algorithm re-written to compute the Montgomery reduction in this new fashion. \begin{figure}[!here] \begin{small} \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{Montgomery Reduction} (modified II). \\ \textbf{Input}. Integer $x$, $n$ and $k$ \\ \textbf{Output}. $\beta^{-k}x \mbox{ (mod }n\mbox{)}$ \\ \hline \\ 1. for $t$ from $0$ to $k - 1$ do \\ \hspace{3mm}1.1 $x \leftarrow x + \mu n \beta^t$ \\ 2. Return $x/\beta^k$. \\ \hline \end{tabular} \end{center} \end{small} \caption{Algorithm Montgomery Reduction (modified II)} \end{figure} The value $\mu n \beta^t$ is a multiple of the modulus $n$ meaning that it will not change the residue. If the first digit of the value $\mu n \beta^t$ equals the negative (modulo $\beta$) of the $t$'th digit of $x$ then the addition will result in a zero digit. This problem breaks down to solving the following congruency. \begin{center} \begin{tabular}{rcl} $x_t + \mu n_0$ & $\equiv$ & $0 \mbox{ (mod }\beta\mbox{)}$ \\ $\mu n_0$ & $\equiv$ & $-x_t \mbox{ (mod }\beta\mbox{)}$ \\ $\mu$ & $\equiv$ & $-x_t/n_0 \mbox{ (mod }\beta\mbox{)}$ \\ \end{tabular} \end{center} In each iteration of the loop on step 1 a new value of $\mu$ must be calculated. The value of $-1/n_0 \mbox{ (mod }\beta\mbox{)}$ is used extensively in this algorithm and should be precomputed. Let $\rho$ represent the negative of the modular inverse of $n_0$ modulo $\beta$. For example, let $\beta = 10$ represent the radix. Let $n = 17$ represent the modulus which implies $k = 2$ and $\rho \equiv 7$. Let $x = 33$ represent the value to reduce. \newpage\begin{figure} \begin{center} \begin{tabular}{|c|c|c|} \hline \textbf{Step ($t$)} & \textbf{Value of $x$} & \textbf{Value of $\mu$} \\ \hline -- & $33$ & --\\ \hline $0$ & $33 + \mu n = 50$ & $1$ \\ \hline $1$ & $50 + \mu n \beta = 900$ & $5$ \\ \hline \end{tabular} \end{center} \caption{Example of Montgomery Reduction} \end{figure} The final result $900$ is then divided by $\beta^k$ to produce the final result $9$. The first observation is that $9 \nequiv x \mbox{ (mod }n\mbox{)}$ which implies the result is not the modular residue of $x$ modulo $n$. However, recall that the residue is actually multiplied by $\beta^{-k}$ in the algorithm. To get the true residue the value must be multiplied by $\beta^k$. In this case $\beta^k \equiv 15 \mbox{ (mod }n\mbox{)}$ and the correct residue is $9 \cdot 15 \equiv 16 \mbox{ (mod }n\mbox{)}$. \subsection{Baseline Montgomery Reduction} The baseline Montgomery reduction algorithm will produce the residue for any size input. It is designed to be a catch-all algororithm for Montgomery reductions. \newpage\begin{figure}[!here] \begin{small} \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{mp\_montgomery\_reduce}. \\ \textbf{Input}. mp\_int $x$, mp\_int $n$ and a digit $\rho \equiv -1/n_0 \mbox{ (mod }n\mbox{)}$. \\ \hspace{11.5mm}($0 \le x < n^2, n > 1, (n, \beta) = 1, \beta^k > n$) \\ \textbf{Output}. $\beta^{-k}x \mbox{ (mod }n\mbox{)}$ \\ \hline \\ 1. $digs \leftarrow 2n.used + 1$ \\ 2. If $digs < MP\_ARRAY$ and $m.used < \delta$ then \\ \hspace{3mm}2.1 Use algorithm fast\_mp\_montgomery\_reduce instead. \\ \\ Setup $x$ for the reduction. \\ 3. If $x.alloc < digs$ then grow $x$ to $digs$ digits. \\ 4. $x.used \leftarrow digs$ \\ \\ Eliminate the lower $k$ digits. \\ 5. For $ix$ from $0$ to $k - 1$ do \\ \hspace{3mm}5.1 $\mu \leftarrow x_{ix} \cdot \rho \mbox{ (mod }\beta\mbox{)}$ \\ \hspace{3mm}5.2 $u \leftarrow 0$ \\ \hspace{3mm}5.3 For $iy$ from $0$ to $k - 1$ do \\ \hspace{6mm}5.3.1 $\hat r \leftarrow \mu n_{iy} + x_{ix + iy} + u$ \\ \hspace{6mm}5.3.2 $x_{ix + iy} \leftarrow \hat r \mbox{ (mod }\beta\mbox{)}$ \\ \hspace{6mm}5.3.3 $u \leftarrow \lfloor \hat r / \beta \rfloor$ \\ \hspace{3mm}5.4 While $u > 0$ do \\ \hspace{6mm}5.4.1 $iy \leftarrow iy + 1$ \\ \hspace{6mm}5.4.2 $x_{ix + iy} \leftarrow x_{ix + iy} + u$ \\ \hspace{6mm}5.4.3 $u \leftarrow \lfloor x_{ix+iy} / \beta \rfloor$ \\ \hspace{6mm}5.4.4 $x_{ix + iy} \leftarrow x_{ix+iy} \mbox{ (mod }\beta\mbox{)}$ \\ \\ Divide by $\beta^k$ and fix up as required. \\ 6. $x \leftarrow \lfloor x / \beta^k \rfloor$ \\ 7. If $x \ge n$ then \\ \hspace{3mm}7.1 $x \leftarrow x - n$ \\ 8. Return(\textit{MP\_OKAY}). \\ \hline \end{tabular} \end{center} \end{small} \caption{Algorithm mp\_montgomery\_reduce} \end{figure} \textbf{Algorithm mp\_montgomery\_reduce.} This algorithm reduces the input $x$ modulo $n$ in place using the Montgomery reduction algorithm. The algorithm is loosely based on algorithm 14.32 of \cite[pp.601]{HAC} except it merges the multiplication of $\mu n \beta^t$ with the addition in the inner loop. The restrictions on this algorithm are fairly easy to adapt to. First $0 \le x < n^2$ bounds the input to numbers in the same range as for the Barrett algorithm. Additionally $n > 1$ will ensure a modular inverse $\rho$ exists. $\rho$ must be calculated in advance of this algorithm. Finally the variable $k$ is fixed and a pseudonym for $n.used$. Step 2 decides whether a faster Montgomery algorithm can be used. It is based on the Comba technique meaning that there are limits on the size of the input. This algorithm is discussed in sub-section 7.3.3. Step 5 is the main reduction loop of the algorithm. The value of $\mu$ is calculated once per iteration in the outer loop. The inner loop calculates $x + \mu n \beta^{ix}$ by multiplying $\mu n$ and adding the result to $x$ shifted by $ix$ digits. Both the addition and multiplication are performed in the same loop to save time and memory. Step 5.4 will handle any additional carries that escape the inner loop. Using a quick inspection this algorithm requires $n$ single precision multiplications for the outer loop and $n^2$ single precision multiplications in the inner loop. In total $n^2 + n$ single precision multiplications which compares favourably to Barrett at $n^2 + 2n - 1$ single precision multiplications. \index{bn\_mp\_montgomery\_reduce.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_mp\_montgomery\_reduce.c \vspace{-3mm} \begin{alltt} 016 017 /* computes xR**-1 == x (mod N) via Montgomery Reduction */ 018 int 019 mp_montgomery_reduce (mp_int * x, mp_int * n, mp_digit rho) 020 \{ 021 int ix, res, digs; 022 mp_digit mu; 023 024 /* can the fast reduction [comba] method be used? 025 * 026 * Note that unlike in mp_mul you're safely allowed *less* 027 * than the available columns [255 per default] since carries 028 * are fixed up in the inner loop. 029 */ 030 digs = n->used * 2 + 1; 031 if ((digs < MP_WARRAY) && 032 n->used < 033 (1 << ((CHAR_BIT * sizeof (mp_word)) - (2 * DIGIT_BIT)))) \{ 034 return fast_mp_montgomery_reduce (x, n, rho); 035 \} 036 037 /* grow the input as required */ 038 if (x->alloc < digs) \{ 039 if ((res = mp_grow (x, digs)) != MP_OKAY) \{ 040 return res; 041 \} 042 \} 043 x->used = digs; 044 045 for (ix = 0; ix < n->used; ix++) \{ 046 /* mu = ai * m' mod b */ 047 mu = (x->dp[ix] * rho) & MP_MASK; 048 049 /* a = a + mu * m * b**i */ 050 \{ 051 register int iy; 052 register mp_digit *tmpn, *tmpx, u; 053 register mp_word r; 054 055 /* aliases */ 056 tmpn = n->dp; 057 tmpx = x->dp + ix; 058 059 /* set the carry to zero */ 060 u = 0; 061 062 /* Multiply and add in place */ 063 for (iy = 0; iy < n->used; iy++) \{ 064 r = ((mp_word) mu) * ((mp_word) * tmpn++) + 065 ((mp_word) u) + ((mp_word) * tmpx); 066 u = (r >> ((mp_word) DIGIT_BIT)); 067 *tmpx++ = (r & ((mp_word) MP_MASK)); 068 \} 069 /* propagate carries */ 070 while (u) \{ 071 *tmpx += u; 072 u = *tmpx >> DIGIT_BIT; 073 *tmpx++ &= MP_MASK; 074 \} 075 \} 076 \} 077 078 /* x = x/b**n.used */ 079 mp_rshd (x, n->used); 080 081 /* if A >= m then A = A - m */ 082 if (mp_cmp_mag (x, n) != MP_LT) \{ 083 return s_mp_sub (x, n, x); 084 \} 085 086 return MP_OKAY; 087 \} \end{alltt} \end{small} This is the baseline implementation of the Montgomery reduction algorithm. Lines 30 to 35 determine if the Comba based routine can be used instead. Line 47 computes the value of $\mu$ for that particular iteration of the outer loop. The multiplication $\mu n \beta^{ix}$ is performed in one step in the inner loop. The alias $tmpx$ refers to the $ix$'th digit of $x$ and the alias $tmpn$ refers to the modulus $n$. \subsection{Faster ``Comba'' Montgomery Reduction} The Montgomery reduction requires fewer single precision multiplications than a Barrett reduction, however it is much slower due to the serial nature of the inner loop. The Barrett reduction algorithm requires two slightly modified multipliers which can be implemented with the Comba technique. The Montgomery reduction algorithm cannot directly use the Comba technique to any significant advantage since the inner loop calculates a $k \times 1$ product $k$ times. The biggest obstacle is that at the $ix$'th iteration of the outer loop the value of $x_{ix}$ is required to calculate $\mu$. This means the carries from $0$ to $ix - 1$ must have been propagated upwards to form a valid $ix$'th digit. The solution as it turns out is very simple. Perform a Comba like multiplier and inside the outer loop just after the inner loop fix up the $ix + 1$'th digit by forwarding the carry. With this change in place the Montgomery reduction algorithm can be performed with a Comba style multiplication loop which substantially increases the speed of the algorithm. \newpage\begin{figure}[!here] \begin{small} \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{fast\_mp\_montgomery\_reduce}. \\ \textbf{Input}. mp\_int $x$, mp\_int $n$ and a digit $\rho \equiv -1/n_0 \mbox{ (mod }n\mbox{)}$. \\ \hspace{11.5mm}($0 \le x < n^2, n > 1, (n, \beta) = 1, \beta^k > n$) \\ \textbf{Output}. $\beta^{-k}x \mbox{ (mod }n\mbox{)}$ \\ \hline \\ Place an array of \textbf{MP\_WARRAY} mp\_word variables called $\hat W$ on the stack. \\ 1. if $x.alloc < n.used + 1$ then grow $x$ to $n.used + 1$ digits. \\ Copy the digits of $x$ into the array $\hat W$ \\ 2. For $ix$ from $0$ to $x.used - 1$ do \\ \hspace{3mm}2.1 $\hat W_{ix} \leftarrow x_{ix}$ \\ 3. For $ix$ from $x.used$ to $2n.used - 1$ do \\ \hspace{3mm}3.1 $\hat W_{ix} \leftarrow 0$ \\ Elimiate the lower $k$ digits. \\ 4. for $ix$ from $0$ to $n.used - 1$ do \\ \hspace{3mm}4.1 $\mu \leftarrow \hat W_{ix} \cdot \rho \mbox{ (mod }\beta\mbox{)}$ \\ \hspace{3mm}4.2 For $iy$ from $0$ to $n.used - 1$ do \\ \hspace{6mm}4.2.1 $\hat W_{iy + ix} \leftarrow \hat W_{iy + ix} + \mu \cdot n_{iy}$ \\ \hspace{3mm}4.3 $\hat W_{ix + 1} \leftarrow \hat W_{ix + 1} + \lfloor \hat W_{ix} / \beta \rfloor$ \\ Propagate carries upwards. \\ 5. for $ix$ from $n.used$ to $2n.used + 1$ do \\ \hspace{3mm}5.1 $\hat W_{ix + 1} \leftarrow \hat W_{ix + 1} + \lfloor \hat W_{ix} / \beta \rfloor$ \\ Shift right and reduce modulo $\beta$ simultaneously. \\ 6. for $ix$ from $0$ to $n.used + 1$ do \\ \hspace{3mm}6.1 $x_{ix} \leftarrow \hat W_{ix + n.used} \mbox{ (mod }\beta\mbox{)}$ \\ Zero excess digits and fixup $x$. \\ 7. if $x.used > n.used + 1$ then do \\ \hspace{3mm}7.1 for $ix$ from $n.used + 1$ to $x.used - 1$ do \\ \hspace{6mm}7.1.1 $x_{ix} \leftarrow 0$ \\ 8. $x.used \leftarrow n.used + 1$ \\ 9. Clamp excessive digits of $x$. \\ 10. If $x \ge n$ then \\ \hspace{3mm}10.1 $x \leftarrow x - n$ \\ 11. Return(\textit{MP\_OKAY}). \\ \hline \end{tabular} \end{center} \end{small} \caption{Algorithm fast\_mp\_montgomery\_reduce} \end{figure} \textbf{Algorithm fast\_mp\_montgomery\_reduce.} This algorithm will compute the Montgomery reduction of $x$ modulo $n$ using the Comba technique. It is on most computer platforms significantly faster than algorithm mp\_montgomery\_reduce and algorithm mp\_reduce (\textit{Barrett reduction}). The algorithm has the same restrictions on the input as the baseline reduction algorithm. An additional two restrictions are imposed on this algorithm. The number of digits $k$ in the the modulus $n$ must not violate $MP\_WARRAY > 2k +1$ and $n < \delta$. When $\beta = 2^{28}$ this algorithm can be used to reduce modulo a modulus of at most $3,556$ bits in length. As in the other Comba reduction algorithms there is a $\hat W$ array which stores the columns of the product. It is initially filled with the contents of $x$ with the excess digits zeroed. The reduction loop is very similar the to the baseline loop at heart. The multiplication on step 4.1 can be single precision only since $ab \mbox{ (mod }\beta\mbox{)} \equiv (a \mbox{ mod }\beta)(b \mbox{ mod }\beta)$. Some multipliers such as those on the ARM processors take a variable length time to complete depending on the number of bytes of result it must produce. By performing a single precision multiplication instead half the amount of time is spent. Also note that digit $\hat W_{ix}$ must have the carry from the $ix - 1$'th digit propagated upwards in order for this to work. That is what step 4.3 will do. In effect over the $n.used$ iterations of the outer loop the $n.used$'th lower columns all have the their carries propagated forwards. Note how the upper bits of those same words are not reduced modulo $\beta$. This is because those values will be discarded shortly and there is no point. Step 5 will propgate the remainder of the carries upwards. On step 6 the columns are reduced modulo $\beta$ and shifted simultaneously as they are stored in the destination $x$. \index{bn\_fast\_mp\_montgomery\_reduce.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_fast\_mp\_montgomery\_reduce.c \vspace{-3mm} \begin{alltt} 016 017 /* computes xR**-1 == x (mod N) via Montgomery Reduction 018 * 019 * This is an optimized implementation of mp_montgomery_reduce 020 * which uses the comba method to quickly calculate the columns of the 021 * reduction. 022 * 023 * Based on Algorithm 14.32 on pp.601 of HAC. 024 */ 025 int 026 fast_mp_montgomery_reduce (mp_int * x, mp_int * n, mp_digit rho) 027 \{ 028 int ix, res, olduse; 029 mp_word W[MP_WARRAY]; 030 031 /* get old used count */ 032 olduse = x->used; 033 034 /* grow a as required */ 035 if (x->alloc < n->used + 1) \{ 036 if ((res = mp_grow (x, n->used + 1)) != MP_OKAY) \{ 037 return res; 038 \} 039 \} 040 041 \{ 042 register mp_word *_W; 043 register mp_digit *tmpx; 044 045 _W = W; 046 tmpx = x->dp; 047 048 /* copy the digits of a into W[0..a->used-1] */ 049 for (ix = 0; ix < x->used; ix++) \{ 050 *_W++ = *tmpx++; 051 \} 052 053 /* zero the high words of W[a->used..m->used*2] */ 054 for (; ix < n->used * 2 + 1; ix++) \{ 055 *_W++ = 0; 056 \} 057 \} 058 059 for (ix = 0; ix < n->used; ix++) \{ 060 /* mu = ai * m' mod b 061 * 062 * We avoid a double precision multiplication (which isn't required) 063 * by casting the value down to a mp_digit. Note this requires 064 * that W[ix-1] have the carry cleared (see after the inner loop) 065 */ 066 register mp_digit mu; 067 mu = (((mp_digit) (W[ix] & MP_MASK)) * rho) & MP_MASK; 068 069 /* a = a + mu * m * b**i 070 * 071 * This is computed in place and on the fly. The multiplication 072 * by b**i is handled by offseting which columns the results 073 * are added to. 074 * 075 * Note the comba method normally doesn't handle carries in the 076 * inner loop In this case we fix the carry from the previous 077 * column since the Montgomery reduction requires digits of the 078 * result (so far) [see above] to work. This is 079 * handled by fixing up one carry after the inner loop. The 080 * carry fixups are done in order so after these loops the 081 * first m->used words of W[] have the carries fixed 082 */ 083 \{ 084 register int iy; 085 register mp_digit *tmpn; 086 register mp_word *_W; 087 088 /* alias for the digits of the modulus */ 089 tmpn = n->dp; 090 091 /* Alias for the columns set by an offset of ix */ 092 _W = W + ix; 093 094 /* inner loop */ 095 for (iy = 0; iy < n->used; iy++) \{ 096 *_W++ += ((mp_word) mu) * ((mp_word) * tmpn++); 097 \} 098 \} 099 100 /* now fix carry for next digit, W[ix+1] */ 101 W[ix + 1] += W[ix] >> ((mp_word) DIGIT_BIT); 102 \} 103 104 105 \{ 106 register mp_digit *tmpx; 107 register mp_word *_W, *_W1; 108 109 /* nox fix rest of carries */ 110 _W1 = W + ix; 111 _W = W + ++ix; 112 113 for (; ix <= n->used * 2 + 1; ix++) \{ 114 *_W++ += *_W1++ >> ((mp_word) DIGIT_BIT); 115 \} 116 117 /* copy out, A = A/b**n 118 * 119 * The result is A/b**n but instead of converting from an 120 * array of mp_word to mp_digit than calling mp_rshd 121 * we just copy them in the right order 122 */ 123 tmpx = x->dp; 124 _W = W + n->used; 125 126 for (ix = 0; ix < n->used + 1; ix++) \{ 127 *tmpx++ = *_W++ & ((mp_word) MP_MASK); 128 \} 129 130 /* zero oldused digits, if the input a was larger than 131 * m->used+1 we'll have to clear the digits */ 132 for (; ix < olduse; ix++) \{ 133 *tmpx++ = 0; 134 \} 135 \} 136 137 /* set the max used and clamp */ 138 x->used = n->used + 1; 139 mp_clamp (x); 140 141 /* if A >= m then A = A - m */ 142 if (mp_cmp_mag (x, n) != MP_LT) \{ 143 return s_mp_sub (x, n, x); 144 \} 145 return MP_OKAY; 146 \} \end{alltt} \end{small} The $\hat W$ array is first filled with digits of $x$ on line 49 then the rest of the digits are zeroed on line 54. Both loops share the same alias variables to make the code easier to read. The value of $\mu$ is calculated in an interesting fashion. First the value $\hat W_{ix}$ is reduced modulo $\beta$ and cast to a mp\_digit. This forces the compiler to use a single precision multiplication and prevents any concerns about loss of precision. Line 101 fixes the carry for the next iteration of the loop by propagating the carry from $\hat W_{ix}$ to $\hat W_{ix+1}$. The for loop on line 113 propagates the rest of the carries upwards through the columns. The for loop on line 126 reduces the columns modulo $\beta$ and shifts them $k$ places at the same time. The alias $\_ \hat W$ actually refers to the array $\hat W$ starting at the $n.used$'th digit, that is $\_ \hat W_{t} = \hat W_{n.used + t}$. \subsection{Montgomery Setup} To calculate the variable $\rho$ a relatively simple algorithm will be required. \begin{figure}[!here] \begin{small} \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{mp\_montgomery\_setup}. \\ \textbf{Input}. mp\_int $n$ ($n > 1$ and $(n, 2) = 1$) \\ \textbf{Output}. $\rho \equiv -1/n_0 \mbox{ (mod }\beta\mbox{)}$ \\ \hline \\ 1. $b \leftarrow n_0$ \\ 2. If $b$ is even return(\textit{MP\_VAL}) \\ 3. $x \leftarrow ((b + 2) \mbox{ AND } 4) << 1) + b$ \\ 4. for $k$ from 0 to $3$ do \\ \hspace{3mm}4.1 $x \leftarrow x \cdot (2 - bx)$ \\ 5. $\rho \leftarrow \beta - x \mbox{ (mod }\beta\mbox{)}$ \\ 6. Return(\textit{MP\_OKAY}). \\ \hline \end{tabular} \end{center} \end{small} \caption{Algorithm mp\_montgomery\_setup} \end{figure} \textbf{Algorithm mp\_montgomery\_setup.} This algorithm will calculate the value of $\rho$ required within the Montgomery reduction algorithms. It uses a very interesting trick to calculate $1/n_0$ when $\beta$ is a power of two. \index{bn\_mp\_montgomery\_setup.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_mp\_montgomery\_setup.c \vspace{-3mm} \begin{alltt} 016 017 /* setups the montgomery reduction stuff */ 018 int 019 mp_montgomery_setup (mp_int * n, mp_digit * rho) 020 \{ 021 mp_digit x, b; 022 023 /* fast inversion mod 2**k 024 * 025 * Based on the fact that 026 * 027 * XA = 1 (mod 2**n) => (X(2-XA)) A = 1 (mod 2**2n) 028 * => 2*X*A - X*X*A*A = 1 029 * => 2*(1) - (1) = 1 030 */ 031 b = n->dp[0]; 032 033 if ((b & 1) == 0) \{ 034 return MP_VAL; 035 \} 036 037 x = (((b + 2) & 4) << 1) + b; /* here x*a==1 mod 2**4 */ 038 x *= 2 - b * x; /* here x*a==1 mod 2**8 */ 039 #if !defined(MP_8BIT) 040 x *= 2 - b * x; /* here x*a==1 mod 2**16 */ 041 #endif 042 #if defined(MP_64BIT) || !(defined(MP_8BIT) || defined(MP_16BIT)) 043 x *= 2 - b * x; /* here x*a==1 mod 2**32 */ 044 #endif 045 #ifdef MP_64BIT 046 x *= 2 - b * x; /* here x*a==1 mod 2**64 */ 047 #endif 048 049 /* rho = -1/m mod b */ 050 *rho = (((mp_digit) 1 << ((mp_digit) DIGIT_BIT)) - x) & MP_MASK; 051 052 return MP_OKAY; 053 \} \end{alltt} \end{small} This source code computes the value of $\rho$ required to perform Montgomery reduction. It has been modified to avoid performing excess multiplications when $\beta$ is not the default 28-bits. \section{The Diminished Radix Algorithm} The diminished radix method of modular reduction \cite{DRMET} is a fairly clever technique which is more efficient than either the Barrett or Montgomery methods. The technique is based on a simple congruence. \begin{equation} (x \mbox{ mod } n) + k \lfloor x / n \rfloor \equiv x \mbox{ (mod }(n - k)\mbox{)} \end{equation} This observation was used in the MMB \cite{MMB} block cipher to create a diffusion primitive. It used the fact that if $n = 2^{31}$ and $k=1$ that then a x86 multiplier could produce the 62-bit product and use the ``shrd'' instruction to perform a double-precision right shift. The proof of the above equation is very simple. First write $x$ in the product form. \begin{equation} x = qn + r \end{equation} Now reduce both sides modulo $(n - k)$. \begin{equation} x \equiv qk + r \mbox{ (mod }(n-k)\mbox{)} \end{equation} The variable $n$ reduces as $n \mbox{ mod } (n - k)$ to $k$. By putting $q = \lfloor x/n \rfloor$ and $r = x \mbox{ mod } n$ into the equation the original congruence is reproduced. The following algorithm is based on these observations. \begin{figure}[!here] \begin{small} \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{Diminished Radix Reduction}. \\ \textbf{Input}. Integer $x$, $n$, $k$ \\ \textbf{Output}. $x \mbox{ mod } (n - k)$ \\ \hline \\ 1. $q \leftarrow \lfloor x / n \rfloor$ \\ 2. $q \leftarrow k \cdot q$ \\ 3. $x \leftarrow x \mbox{ (mod }n\mbox{)}$ \\ 4. $x \leftarrow x + q$ \\ 5. If $x \ge (n - k)$ then \\ \hspace{3mm}5.1 $x \leftarrow x - (n - k)$ \\ \hspace{3mm}5.2 Goto step 1. \\ 6. Return $x$ \\ \hline \end{tabular} \end{center} \end{small} \caption{Algorithm Diminished Radix Reduction} \label{fig:DR} \end{figure} This algorithm will reduce $x$ modulo $n - k$ and return the residue. If $0 \le x < (n - k)^2$ then the algorithm will loop almost always once or twice and occasionally three times. For simplicity sake the value of $x$ is bounded by the following simple polynomial. \begin{equation} 0 \le x < n^2 + k^2 - 2nk \end{equation} The true bound is $0 \le x < (n - k - 1)^2$ but this has quite a few more terms. The value of $q$ after step 1 is bounded by the following. \begin{equation} q < n - 2k - k^2/n \end{equation} Since $k^2$ is going to be considerably smaller than $n$ that term will always be zero. The value of $x$ after step 3 is bounded trivially as $0 \le x < n$. By step four the sum $x + q$ is bounded by \begin{equation} 0 \le q + x < (k + 1)n - 2k^2 - 1 \end{equation} As a result at most $k$ subtractions of $n$ are required to produce the residue. With a second pass $q$ will be loosely bounded by $0 \le q < k^2$ after step 2 while $x$ will still be loosely bounded by $0 \le x < n$ after step 3. After the second pass it is highly unlike that the sum in step 4 will exceed $n - k$. In practice fewer than three passes of the algorithm are required to reduce virtually every input in the range $0 \le x < (n - k - 1)^2$. \subsection{Choice of Moduli} On the surface this algorithm looks like a very expensive algorithm. It requires a couple of subtractions followed by multiplication and other modular reductions. The usefulness of this algorithm becomes exceedingly clear when an appropriate moduli is chosen. Division in general is a very expensive operation to perform. The one exception is when the division is by a power of the radix of representation used. Division by ten for example is simple for humans since it amounts to shifting the decimal place. Similarly division by two (\textit{or powers of two}) is very simple for computers to perform. It would therefore seem logical to choose $n$ of the form $2^p$ which would imply that $\lfloor x / n \rfloor$ is a simple shift of $x$ right $p$ bits. However, there is one operation related to division of power of twos that is even faster than this. If $n = \beta^p$ then the division may be performed by moving whole digits to the right $p$ places. In practice division by $\beta^p$ is much faster than division by $2^p$ for any $p$. Also with the choice of $n = \beta^p$ reducing $x$ modulo $n$ requires zeroing the digits above the $p-1$'th digit of $x$. Throughout the next section the term ``restricted modulus'' will refer to a modulus of the form $\beta^p - k$ where as the term ``unrestricted modulus'' will refer to a modulus of the form $2^p - k$. The word ``restricted'' in this case refers to the fact that it is based on the $2^p$ logic except $p$ must be a multiple of $lg(\beta)$. \subsection{Choice of $k$} Now that division and reduction (\textit{step 1 and 3 of figure~\ref{fig:DR}}) have been optimized to simple digit operations the multiplication by $k$ in step 2 is the most expensive operation. Fortunately the choice of $k$ is not terribly limited. For all intents and purposes it might as well be a single digit. \subsection{Restricted Diminished Radix Reduction} The restricted Diminished Radix algorithm can quickly reduce numbers modulo numbers of the form $n = \beta^p - k$. This algorithm can reduce an input $x$ within the range $0 \le x < n^2$ using a couple passes of the algorithm demonstrated in figure~\ref{fig:DR}. The implementation of this algorithm has been optimized to avoid additional overhead associated with a division by $\beta^p$, the multiplication by $k$ or the addition of $x$ and $q$. The resulting algorithm is very efficient and can lead to substantial improvements when modular exponentiations are performed compared to Montgomery based reduction algorithms. \begin{figure}[!here] \begin{small} \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{mp\_dr\_reduce}. \\ \textbf{Input}. mp\_int $x$, $n$ and a mp\_digit $k = \beta - n_0$ \\ \hspace{11.5mm}($0 \le x < n^2$, $n > 1$, $0 < k \le \beta$) \\ \textbf{Output}. $x \mbox{ mod } n$ \\ \hline \\ 1. $m \leftarrow n.used$ \\ 2. If $x.alloc < 2m$ then grow $x$ to $2m$ digits. \\ 3. $\mu \leftarrow 0$ \\ 4. for $i$ from $0$ to $m - 1$ do \\ \hspace{3mm}4.1 $\hat r \leftarrow k \cdot x_{m+i} + x_{i} + \mu$ \\ \hspace{3mm}4.2 $x_{i} \leftarrow \hat r \mbox{ (mod }\beta\mbox{)}$ \\ \hspace{3mm}4.3 $\mu \leftarrow \lfloor \hat r / \beta \rfloor$ \\ 5. $x_{m} \leftarrow \mu$ \\ 6. for $i$ from $m + 1$ to $x.used - 1$ do \\ \hspace{3mm}6.1 $x_{i} \leftarrow 0$ \\ 7. Clamp excess digits of $x$. \\ 8. If $x \ge n$ then \\ \hspace{3mm}8.1 $x \leftarrow x - n$ \\ \hspace{3mm}8.2 Goto step 3. \\ 9. Return(\textit{MP\_OKAY}). \\ \hline \end{tabular} \end{center} \end{small} \caption{Algorithm mp\_dr\_reduce} \end{figure} \textbf{Algorithm mp\_dr\_reduce.} This algorithm will perform the dimished radix reduction of $x$ modulo $n$. It has similar restrictions to that of the Barrett reduction with the addition that $n$ must be of the form $n = \beta^m - k$ where $0 < k \le \beta$. This algorithm essentially implements the pseudo-code in figure 7.10 except with a slight optimization. The division by $\beta^m$, multiplication by $k$ and addition of $x \mbox{ mod }\beta^m$ are all performed as one step inside the loop on step 4. The division by $\beta^m$ is emulated by accessing the term at the $m+i$'th position which is subsequently multiplied by $k$ and added to the term at the $i$'th position. After the loop the $m$'th digit is set to the carry and the upper digits are zeroed. Step 5 and 6 emulate the reduction modulo $\beta^m$ that should have happend to $x$ before the addition of the multiple of the upper half. At step 8 if $x$ is still larger than $n$ another pass of the algorithm is required. First $n$ is subtracted from $x$ and then the algorithm resumes at step 3. \index{bn\_mp\_dr\_reduce.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_mp\_dr\_reduce.c \vspace{-3mm} \begin{alltt} 016 017 /* reduce "x" in place modulo "n" using the Diminished Radix algorithm. 018 * 019 * Based on algorithm from the paper 020 * 021 * "Generating Efficient Primes for Discrete Log Cryptosystems" 022 * Chae Hoon Lim, Pil Loong Lee, 023 * POSTECH Information Research Laboratories 024 * 025 * The modulus must be of a special format [see manual] 026 * 027 * Has been modified to use algorithm 7.10 from the LTM book instead 028 */ 029 int 030 mp_dr_reduce (mp_int * x, mp_int * n, mp_digit k) 031 \{ 032 int err, i, m; 033 mp_word r; 034 mp_digit mu, *tmpx1, *tmpx2; 035 036 /* m = digits in modulus */ 037 m = n->used; 038 039 /* ensure that "x" has at least 2m digits */ 040 if (x->alloc < m + m) \{ 041 if ((err = mp_grow (x, m + m)) != MP_OKAY) \{ 042 return err; 043 \} 044 \} 045 046 /* top of loop, this is where the code resumes if 047 * another reduction pass is required. 048 */ 049 top: 050 /* aliases for digits */ 051 /* alias for lower half of x */ 052 tmpx1 = x->dp; 053 054 /* alias for upper half of x, or x/B**m */ 055 tmpx2 = x->dp + m; 056 057 /* set carry to zero */ 058 mu = 0; 059 060 /* compute (x mod B**m) + mp * [x/B**m] inline and inplace */ 061 for (i = 0; i < m; i++) \{ 062 r = ((mp_word)*tmpx2++) * ((mp_word)k) + *tmpx1 + mu; 063 *tmpx1++ = r & MP_MASK; 064 mu = r >> ((mp_word)DIGIT_BIT); 065 \} 066 067 /* set final carry */ 068 *tmpx1++ = mu; 069 070 /* zero words above m */ 071 for (i = m + 1; i < x->used; i++) \{ 072 *tmpx1++ = 0; 073 \} 074 075 /* clamp, sub and return */ 076 mp_clamp (x); 077 078 /* if x >= n then subtract and reduce again 079 * Each successive "recursion" makes the input smaller and smaller. 080 */ 081 if (mp_cmp_mag (x, n) != MP_LT) \{ 082 s_mp_sub(x, n, x); 083 goto top; 084 \} 085 return MP_OKAY; 086 \} \end{alltt} \end{small} The first step is to grow $x$ as required to $2m$ digits since the reduction is performed in place on $x$. The label on line 49 is where the algorithm will resume if further reduction passes are required. In theory it could be placed at the top of the function however, the size of the modulus and question of whether $x$ is large enough are invariant after the first pass meaning that it would be a waste of time. The aliases $tmpx1$ and $tmpx2$ refer to the digits of $x$ where the latter is offset by $m$ digits. By reading digits from $x$ offset by $m$ digits a division by $\beta^m$ can be simulated virtually for free. The loop on line 61 performs the bulk of the work (\textit{corresponds to step 4 of algorithm 7.11}) in this algorithm. By line 68 the pointer $tmpx1$ points to the $m$'th digit of $x$ which is where the final carry will be placed. Similarly by line 71 the same pointer will point to the $m+1$'th digit where the zeroes will be placed. Since the algorithm is only valid if both $x$ and $n$ are greater than zero an unsigned comparison suffices to determine if another pass is required. With the same logic at line 82 the value of $x$ is known to be greater than or equal to $n$ meaning that an unsigned subtraction can be used as well. Since the destination of the subtraction is the larger of the inputs the call to algorithm s\_mp\_sub cannot fail and the return code does not need to be checked. \subsubsection{Setup} To setup the restricted Diminished Radix algorithm the value $k = \beta - n_0$ is required. This algorithm is not really complicated but provided for completeness. \begin{figure}[!here] \begin{small} \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{mp\_dr\_setup}. \\ \textbf{Input}. mp\_int $n$ \\ \textbf{Output}. $k = \beta - n_0$ \\ \hline \\ 1. $k \leftarrow \beta - n_0$ \\ \hline \end{tabular} \end{center} \end{small} \caption{Algorithm mp\_dr\_setup} \end{figure} \index{bn\_mp\_dr\_setup.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_mp\_dr\_setup.c \vspace{-3mm} \begin{alltt} 016 017 /* determines the setup value */ 018 void mp_dr_setup(mp_int *a, mp_digit *d) 019 \{ 020 /* the casts are required if DIGIT_BIT is one less than 021 * the number of bits in a mp_digit [e.g. DIGIT_BIT==31] 022 */ 023 *d = (mp_digit)((((mp_word)1) << ((mp_word)DIGIT_BIT)) - 024 ((mp_word)a->dp[0])); 025 \} 026 \end{alltt} \end{small} \subsubsection{Modulus Detection} Another algorithm which will be useful is the ability to detect a restricted Diminished Radix modulus. An integer is said to be of restricted Diminished Radix form if all of the digits are equal to $\beta - 1$ except the trailing digit which may be any value. \begin{figure}[!here] \begin{small} \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{mp\_dr\_is\_modulus}. \\ \textbf{Input}. mp\_int $n$ \\ \textbf{Output}. $1$ if $n$ is in D.R form, $0$ otherwise \\ \hline 1. If $n.used < 2$ then return($0$). \\ 2. for $ix$ from $1$ to $n.used - 1$ do \\ \hspace{3mm}2.1 If $n_{ix} \ne \beta - 1$ return($0$). \\ 3. Return($1$). \\ \hline \end{tabular} \end{center} \end{small} \caption{Algorithm mp\_dr\_is\_modulus} \end{figure} \textbf{Algorithm mp\_dr\_is\_modulus.} This algorithm determines if a value is in Diminished Radix form. Step 1 rejects obvious cases where fewer than two digits are in the mp\_int. Step 2 tests all but the first digit to see if they are equal to $\beta - 1$. If the algorithm manages to get to step 3 then $n$ must of Diminished Radix form. \index{bn\_mp\_dr\_is\_modulus.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_mp\_dr\_is\_modulus.c \vspace{-3mm} \begin{alltt} 016 017 /* determines if a number is a valid DR modulus */ 018 int mp_dr_is_modulus(mp_int *a) 019 \{ 020 int ix; 021 022 /* must be at least two digits */ 023 if (a->used < 2) \{ 024 return 0; 025 \} 026 027 for (ix = 1; ix < a->used; ix++) \{ 028 if (a->dp[ix] != MP_MASK) \{ 029 return 0; 030 \} 031 \} 032 return 1; 033 \} 034 \end{alltt} \end{small} \subsection{Unrestricted Diminished Radix Reduction} The unrestricted Diminished Radix algorithm allows modular reductions to be performed when the modulus is of the form $2^p - k$. This algorithm is a straightforward adaptation of algorithm~\ref{fig:DR}. In general the restricted Diminished Radix reduction algorithm is much faster since it has considerably lower overhead. However, this new algorithm is much faster than either Montgomery or Barrett reduction when the moduli are of the appropriate form. \begin{figure}[!here] \begin{small} \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{mp\_reduce\_2k}. \\ \textbf{Input}. mp\_int $a$ and $n$. mp\_digit $k$ \\ \hspace{11.5mm}($a \ge 0$, $n > 1$, $0 < k < \beta$, $n + k$ is a power of two) \\ \textbf{Output}. $a \mbox{ (mod }n\mbox{)}$ \\ \hline 1. $p \leftarrow \lfloor lg(n) \rfloor + 1$ (\textit{mp\_count\_bits}) \\ 2. While $a \ge n$ do \\ \hspace{3mm}2.1 $q \leftarrow \lfloor a / 2^p \rfloor$ (\textit{mp\_div\_2d}) \\ \hspace{3mm}2.2 $a \leftarrow a \mbox{ (mod }2^p\mbox{)}$ (\textit{mp\_mod\_2d}) \\ \hspace{3mm}2.3 $q \leftarrow q \cdot k$ (\textit{mp\_mul\_d}) \\ \hspace{3mm}2.4 $a \leftarrow a - q$ (\textit{s\_mp\_sub}) \\ \hspace{3mm}2.5 If $a \ge n$ then do \\ \hspace{6mm}2.5.1 $a \leftarrow a - n$ \\ 3. Return(\textit{MP\_OKAY}). \\ \hline \end{tabular} \end{center} \end{small} \caption{Algorithm mp\_reduce\_2k} \end{figure} \textbf{Algorithm mp\_reduce\_2k.} This algorithm quickly reduces an input $a$ modulo an unrestricted Diminished Radix modulus $n$. \index{bn\_mp\_reduce\_2k.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_mp\_reduce\_2k.c \vspace{-3mm} \begin{alltt} 016 017 /* reduces a modulo n where n is of the form 2**p - k */ 018 int 019 mp_reduce_2k(mp_int *a, mp_int *n, mp_digit k) 020 \{ 021 mp_int q; 022 int p, res; 023 024 if ((res = mp_init(&q)) != MP_OKAY) \{ 025 return res; 026 \} 027 028 p = mp_count_bits(n); 029 top: 030 /* q = a/2**p, a = a mod 2**p */ 031 if ((res = mp_div_2d(a, p, &q, a)) != MP_OKAY) \{ 032 goto ERR; 033 \} 034 035 if (k != 1) \{ 036 /* q = q * k */ 037 if ((res = mp_mul_d(&q, k, &q)) != MP_OKAY) \{ 038 goto ERR; 039 \} 040 \} 041 042 /* a = a + q */ 043 if ((res = s_mp_add(a, &q, a)) != MP_OKAY) \{ 044 goto ERR; 045 \} 046 047 if (mp_cmp_mag(a, n) != MP_LT) \{ 048 s_mp_sub(a, n, a); 049 goto top; 050 \} 051 052 ERR: 053 mp_clear(&q); 054 return res; 055 \} 056 \end{alltt} \end{small} \subsubsection{Unrestricted Setup} To setup this reduction algorithm the value of $k = 2^p - n$ is required. \begin{figure}[!here] \begin{small} \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{mp\_reduce\_2k\_setup}. \\ \textbf{Input}. mp\_int $n$ \\ \textbf{Output}. $k = 2^p - n$ \\ \hline 1. $p \leftarrow \lfloor lg(n) \rfloor + 1$ (\textit{mp\_count\_bits}) \\ 2. $x \leftarrow 2^p$ (\textit{mp\_2expt}) \\ 3. $x \leftarrow x - n$ (\textit{mp\_sub}) \\ 4. $k \leftarrow x_0$ \\ 5. Return(\textit{MP\_OKAY}). \\ \hline \end{tabular} \end{center} \end{small} \caption{Algorithm mp\_reduce\_2k\_setup} \end{figure} \textbf{Algorithm mp\_reduce\_2k\_setup.} \index{bn\_mp\_reduce\_2k\_setup.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_mp\_reduce\_2k\_setup.c \vspace{-3mm} \begin{alltt} 016 017 /* determines the setup value */ 018 int 019 mp_reduce_2k_setup(mp_int *a, mp_digit *d) 020 \{ 021 int res, p; 022 mp_int tmp; 023 024 if ((res = mp_init(&tmp)) != MP_OKAY) \{ 025 return res; 026 \} 027 028 p = mp_count_bits(a); 029 if ((res = mp_2expt(&tmp, p)) != MP_OKAY) \{ 030 mp_clear(&tmp); 031 return res; 032 \} 033 034 if ((res = s_mp_sub(&tmp, a, &tmp)) != MP_OKAY) \{ 035 mp_clear(&tmp); 036 return res; 037 \} 038 039 *d = tmp.dp[0]; 040 mp_clear(&tmp); 041 return MP_OKAY; 042 \} \end{alltt} \end{small} \subsubsection{Unrestricted Detection} An integer $n$ is a valid unrestricted Diminished Radix modulus if either of the following are true. \begin{enumerate} \item The number has only one digit. \item The number has more than one digit and every bit from the $\beta$'th to the most significant is one. \end{enumerate} If either condition is true than there is a power of two namely $2^p$ such that $0 < 2^p - n < \beta$. -- Finish this section later, Tom. \section{Algorithm Comparison} So far three very different algorithms for modular reduction have been discussed. Each of the algorithms have their own strengths and weaknesses that makes having such a selection very useful. The following table sumarizes the three algorithms along with comparisons of work factors. Since all three algorithms have the restriction that $0 \le x < n^2$ and $n > 1$ those limitations are not included in the table. \begin{center} \begin{small} \begin{tabular}{|c|c|c|c|c|c|} \hline \textbf{Method} & \textbf{Work Required} & \textbf{Limitations} & \textbf{$m = 8$} & \textbf{$m = 32$} & \textbf{$m = 64$} \\ \hline Barrett & $m^2 + 2m - 1$ & None & $79$ & $1087$ & $4223$ \\ \hline Montgomery & $m^2 + m$ & $n$ must be odd & $72$ & $1056$ & $4160$ \\ \hline D.R. & $2m$ & $n = \beta^m - k$ & $16$ & $64$ & $128$ \\ \hline \end{tabular} \end{small} \end{center} In theory Montgomery and Barrett reductions would require roughly the same amount of time to complete. However, in practice since Montgomery reduction can be written as a single function with the Comba technique it is much faster. Barrett reduction suffers from the overhead of calling the half precision multipliers, addition and division by $\beta$ algorithms. For almost every cryptographic algorithm Montgomery reduction is the algorithm of choice. The one set of algorithms where Diminished Radix reduction truly shines are based on the discrete logarithm problem such as Diffie-Hellman \cite{DH} and ElGamal \cite{ELGAMAL}. In these algorithms primes of the form $\beta^m - k$ can be found and shared amongst users. These primes will allow the Diminished Radix algorithm to be used in modular exponentiation to greatly speed up the operation. \section*{Exercises} \begin{tabular}{cl} $\left [ 3 \right ]$ & Prove that the ``trick'' in algorithm mp\_montgomery\_setup actually \\ & calculates the correct value of $\rho$. \\ & \\ $\left [ 2 \right ]$ & Devise an algorithm to reduce modulo $n + k$ for small $k$ quickly. \\ & \\ $\left [ 4 \right ]$ & Prove that the pseudo-code algorithm ``Diminished Radix Reduction'' \\ & (\textit{figure 7.10}) terminates. Also prove the probability that it will \\ & terminate within $1 \le k \le 10$ iterations. \\ & \\ \end{tabular} \chapter{Exponentiation} Exponentiation is the operation of raising one variable to the power of another, for example, $a^b$. A variant of exponentiation, computed in a finite field or ring, is called modular exponentiation. This latter style of operation is typically used in public key cryptosystems such as RSA and Diffie-Hellman. The ability to quickly compute modular exponentiations is of great benefit to any such cryptosystem and many methods have been sought to speed it up. \section{Exponentiation Basics} A trivial algorithm would simply multiply $a$ against itself $b - 1$ times to compute the exponentiation desired. However, as $b$ grows in size the number of multiplications becomes prohibitive. Imagine what would happen if $b$ $\approx$ $2^{1024}$ as is the case when computing an RSA signature with a $1024$-bit key. Such a calculation could never be completed as it would take simply far too long. Fortunately there is a very simple algorithm based on the laws of exponents. Recall that $lg_a(a^b) = b$ and that $lg_a(a^ba^c) = b + c$ which are two trivial relationships between the base and the exponent. Let $b_i$ represent the $i$'th bit of $b$ starting from the least significant bit. If $b$ is a $k$-bit integer than the following equation is true. \begin{equation} a^b = \prod_{i=0}^{k-1} a^{2^i \cdot b_i} \end{equation} By taking the base $a$ logarithm of both sides of the equation the following equation is the result. \begin{equation} b = \sum_{i=0}^{k-1}2^i \cdot b_i \end{equation} This is indeed true. The term $a^{2^i}$ can be found from the $i - 1$'th term by squaring the term since $\left ( a^{2^i} \right )^2$ is equal to $a^{2^{i+1}}$. This trivial algorithm forms the basis of essentially all fast exponentiation algorithms. It requires $k$ squarings and on average $k \over 2$ multiplications to compute the result. This is indeed quite an improvement over simply multiplying by $a$ a total of $b-1$ times. While this current method is a considerable speed up there are further improvements to be made. For example, the $a^{2^i}$ term does not need to be an auxilary variable. Consider the following algorithm. \begin{figure}[!here] \begin{small} \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{Left to Right Exponentiation}. \\ \textbf{Input}. Integer $a$, $b$ and $k$ \\ \textbf{Output}. $c = a^b$ \\ \hline \\ 1. $c \leftarrow 1$ \\ 2. for $i$ from $k - 1$ to $0$ do \\ \hspace{3mm}2.1 $c \leftarrow c^2$ \\ \hspace{3mm}2.2 $c \leftarrow c \cdot a^{b_i}$ \\ 3. Return $c$. \\ \hline \end{tabular} \end{center} \end{small} \caption{Left to Right Exponentiation} \end{figure} This algorithm starts from the most significant bit and works towards the least significant bit. When the $i$'th bit of $b$ is set $a$ is multiplied against the current product. In each iteration the product is squared which doubles the exponent of the individual terms of the product. For example, let $b = 101100_2 \equiv 44_{10}$. The following chart demonstrates the actions of the algorithm. \newpage\begin{figure} \begin{center} \begin{tabular}{|c|c|} \hline \textbf{Value of $i$} & \textbf{Value of $c$} \\ \hline - & $1$ \\ \hline $5$ & $a$ \\ \hline $4$ & $a^2$ \\ \hline $3$ & $a^4 \cdot a$ \\ \hline $2$ & $a^8 \cdot a^2 \cdot a$ \\ \hline $1$ & $a^{16} \cdot a^4 \cdot a^2$ \\ \hline $0$ & $a^{32} \cdot a^8 \cdot a^4$ \\ \hline \end{tabular} \end{center} \caption{Example of Left to Right Exponentiation} \end{figure} When the product $a^{32} \cdot a^8 \cdot a^4$ is simplified it is equal $a^{44}$ which is the desired exponentiation. This particular algorithm is called ``Left to Right'' because it reads the exponent in that order. All of the exponentiation algorithms that will be presented are of this nature. \subsection{Single Digit Exponentiation} The first algorithm in the series of exponentiation algorithms will be an unbounded algorithm where the exponent is a single digit. It is intended to be used when a small power of an input is required (\textit{e.g. $a^5$}). It is faster than simply multiplying $b - 1$ times for all values of $b$ that are greater than three. \newpage\begin{figure}[!here] \begin{small} \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{mp\_expt\_d}. \\ \textbf{Input}. mp\_int $a$ and mp\_digit $b$ \\ \textbf{Output}. $c = a^b$ \\ \hline \\ 1. $g \leftarrow a$ (\textit{mp\_init\_copy}) \\ 2. $c \leftarrow 1$ (\textit{mp\_set}) \\ 3. for $x$ from 0 to $lg(\beta) - 1$ do \\ \hspace{3mm}3.1 $c \leftarrow c^2$ (\textit{mp\_sqr}) \\ \hspace{3mm}3.2 If $b$ AND $2^{lg(\beta) - 1} \ne 0$ then \\ \hspace{6mm}3.2.1 $c \leftarrow c \cdot g$ (\textit{mp\_mul}) \\ \hspace{3mm}3.3 $b \leftarrow b << 1$ \\ 4. Clear $g$. \\ 5. Return(\textit{MP\_OKAY}). \\ \hline \end{tabular} \end{center} \end{small} \caption{Algorithm mp\_expt\_d} \end{figure} \textbf{Algorithm mp\_expt\_d.} This algorithm computes the value of $a$ raised to the power of a single digit $b$. It uses the left to right exponentiation algorithm to quickly compute the exponentiation. It is loosely based on algorithm 14.79 of HAC \cite[pp. 615]{HAC} with the difference that the exponent is a fixed width. A copy of $a$ is made on the first step to allow destination variable $c$ be the same as the source variable $a$. The result is set to the initial value of $1$ in the subsequent step. Inside the loop the exponent is read from the most significant bit first downto the least significant bit. First $c$ is invariably squared on step 3.1. In the following step if the most significant bit of $b$ is one the copy of $a$ is multiplied against the result. The value of $b$ is shifted left one bit to make the next bit down from the most signficant bit become the new most significant bit. In effect each iteration of the loop moves the bits of the exponent $b$ upwards to the most significant location. \index{bn\_mp\_expt\_d.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_mp\_expt\_d.c \vspace{-3mm} \begin{alltt} 016 017 /* calculate c = a**b using a square-multiply algorithm */ 018 int 019 mp_expt_d (mp_int * a, mp_digit b, mp_int * c) 020 \{ 021 int res, x; 022 mp_int g; 023 024 if ((res = mp_init_copy (&g, a)) != MP_OKAY) \{ 025 return res; 026 \} 027 028 /* set initial result */ 029 mp_set (c, 1); 030 031 for (x = 0; x < (int) DIGIT_BIT; x++) \{ 032 /* square */ 033 if ((res = mp_sqr (c, c)) != MP_OKAY) \{ 034 mp_clear (&g); 035 return res; 036 \} 037 038 /* if the bit is set multiply */ 039 if ((b & (mp_digit) (((mp_digit)1) << (DIGIT_BIT - 1))) != 0) \{ 040 if ((res = mp_mul (c, &g, c)) != MP_OKAY) \{ 041 mp_clear (&g); 042 return res; 043 \} 044 \} 045 046 /* shift to next bit */ 047 b <<= 1; 048 \} 049 050 mp_clear (&g); 051 return MP_OKAY; 052 \} \end{alltt} \end{small} -- Some note later. \subsection{$k$-ary Exponentiation} When calculating an exponentiation the most time consuming bottleneck is the multiplications which are in general a small factor slower than squaring. Recall from the previous algorithm that $b_{i}$ refers to the $i$'th bit of the exponent $b$. Suppose it referred to the $i$'th $k$-bit digit of the exponent of $b$. For $k = 1$ the definitions are synonymous and for $k > 1$ the resulting algorithm computes the same exponentiation. A group of $k$ bits from the exponent is called a \textit{window}. That is it is a window on a small portion of the exponent. Consider the following modification to the basic left to right exponentiation algorithm. \newpage\begin{figure}[!here] \begin{small} \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{$k$-ary Exponentiation}. \\ \textbf{Input}. Integer $a$, $b$, $k$ and $t$ \\ \textbf{Output}. $c = a^b$ \\ \hline \\ 1. $c \leftarrow 1$ \\ 2. for $i$ from $t - 1$ to $0$ do \\ \hspace{3mm}2.1 $c \leftarrow c^{2^k} $ \\ \hspace{3mm}2.2 Extract the $i$'th $k$-bit word from $b$ and store it in $g$. \\ \hspace{3mm}2.3 $c \leftarrow c \cdot a^g$ \\ 3. Return $c$. \\ \hline \end{tabular} \end{center} \end{small} \caption{$k$-ary Exponentiation} \end{figure} The squaring on step 2.1 can be calculated by squaring the value $c$ successively $k$ times. If the values of $a^g$ for $0 < g < 2^k$ have been precomputed this algorithm requires only $t$ multiplications and $tk$ squarings. The table can be generated with $2^{k - 1} - 1$ squarings and $2^{k - 1} + 1$ multiplications. This algorithm assumes that the number of bits in the exponent is evenly divisible by $k$. However, when it is not the remaining $0 < x \le k - 1$ bits can be handled with the original left to right style algorithm. Suppose $k = 4$ and $t = 100$. This modified algorithm will require $109$ multiplications and $408$ squarings to compute the exponentiation. The original algorithm would on average have required $200$ multiplications and $400$ squrings to compute the same value. The total number of squarings has increased slightly but the number of multiplications has nearly halved. \subsection{Sliding-Window Exponentiation} A simple modification to the previous algorithm is only generate the upper half of the table in the range $2^{k-1} \le g < 2^k$. Essentially this is a table for all values of $g$ where the most significant bit of $g$ is a one. However, in order for this to be allowed in the algorithm values of $g$ in the range $0 \le g < 2^{k-1}$ must be avoided. \newpage\begin{figure}[!here] \begin{small} \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{Sliding Window $k$-ary Exponentiation}. \\ \textbf{Input}. Integer $a$, $b$, $k$ and $t$ \\ \textbf{Output}. $c = a^b$ \\ \hline \\ 1. $c \leftarrow 1$ \\ 2. for $i$ from $t - 1$ to $0$ do \\ \hspace{3mm}2.1 If the $i$'th bit of $b$ is a zero then \\ \hspace{6mm}2.1.1 $c \leftarrow c^2$ \\ \hspace{3mm}2.2 else do \\ \hspace{6mm}2.2.1 $c \leftarrow c^{2^k}$ \\ \hspace{6mm}2.2.2 Extract the $k$ bits from $(b_{i}b_{i-1}\ldots b_{i-(k-1)})$ and store it in $g$. \\ \hspace{6mm}2.2.3 $c \leftarrow c \cdot a^g$ \\ \hspace{6mm}2.2.4 $i \leftarrow i - k$ \\ 3. Return $c$. \\ \hline \end{tabular} \end{center} \end{small} \caption{Sliding Window $k$-ary Exponentiation} \end{figure} Similar to the previous algorithm this algorithm must have a special handler when fewer than $k$ bits are left in the exponent. While this algorithm requires the same number of squarings it can potentially have fewer multiplications. The pre-computed table $a^g$ is also half the size as the previous table. Consider the exponent $b = 111101011001000_2 \equiv 31432_{10}$ with $k = 3$ using both algorithms. The first algorithm will divide the exponent up as the following five $3$-bit words $b \equiv \left ( 111, 101, 011, 001, 000 \right )_{2}$. The second algorithm will break the exponent as $b \equiv \left ( 111, 101, 0, 110, 0, 100, 0 \right )_{2}$. The single digit $0$ in the second representation are where a single squaring took place instead of a squaring and multiplication. In total the first method requires $10$ multiplications and $18$ squarings. The second method requires $8$ multiplications and $18$ squarings. In general the sliding window method is never slower than the generic $k$-ary method and often it is slightly faster. \section{Modular Exponentiation} Modular exponentiation is essentially computing the power of a base within a finite field or ring. For example, computing $d \equiv a^b \mbox{ (mod }c\mbox{)}$ is a modular exponentiation. Instead of first computing $a^b$ and then reducing it modulo $c$ the intermediate result is reduced modulo $c$ after every squaring or multiplication operation. This guarantees that any intermediate result is bounded by $0 \le d \le c^2 - 2c + 1$ and can be reduced modulo $c$ quickly using any of the three algorithms presented in chapter seven. Before the actual modular exponentiation algorithm can be written a wrapper algorithm must be written first. This wrapper algorithm will allow the exponent $b$ to be negative which is computed as $c \equiv \left (1 / a \right )^{\vert b \vert} \mbox{(mod }d\mbox{)}$. The value of $(1/a) \mbox{ mod }c$ is computed using the modular inverse (\textit{see section 10.4}). If no inverse exists the algorithm terminates with an error. \begin{figure}[!here] \begin{small} \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{mp\_exptmod}. \\ \textbf{Input}. mp\_int $a$, $b$ and $c$ \\ \textbf{Output}. $y \equiv g^x \mbox{ (mod }p\mbox{)}$ \\ \hline \\ 1. If $c.sign = MP\_NEG$ return(\textit{MP\_VAL}). \\ 2. If $b.sign = MP\_NEG$ then \\ \hspace{3mm}2.1 $g' \leftarrow g^{-1} \mbox{ (mod }c\mbox{)}$ \\ \hspace{3mm}2.2 $x' \leftarrow \vert x \vert$ \\ \hspace{3mm}2.3 Compute $d \equiv g'^{x'} \mbox{ (mod }c\mbox{)}$ via recursion. \\ 3. if ($p$ is odd \textbf{OR} $p$ is a D.R. modulus) \textbf{AND} $p.used > 4$ then \\ \hspace{3mm}3.1 Compute $y \equiv g^{x} \mbox{ (mod }p\mbox{)}$ via algorithm mp\_exptmod\_fast. \\ 4. else \\ \hspace{3mm}4.1 Compute $y \equiv g^{x} \mbox{ (mod }p\mbox{)}$ via algorithm s\_mp\_exptmod. \\ \hline \end{tabular} \end{center} \end{small} \caption{Algorithm mp\_exptmod} \end{figure} \textbf{Algorithm mp\_exptmod.} The first algorithm which actually performs modular exponentiation is algorithm s\_mp\_exptmod. It is a sliding window $k$-ary algorithm which uses Barrett reduction to reduce the product modulo $p$. The second algorithm mp\_exptmod\_fast performs the same operation except it uses either Montgomery or Diminished Radix reduction. The two latter reduction algorithms are clumped in the same exponentiation algorithm since their arguments are essentially the same (\textit{two mp\_ints and one mp\_digit}). \index{bn\_mp\_exptmod.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_mp\_exptmod.c \vspace{-3mm} \begin{alltt} 016 017 018 /* this is a shell function that calls either the normal or Montgomery 019 * exptmod functions. Originally the call to the montgomery code was 020 * embedded in the normal function but that wasted alot of stack space 021 * for nothing (since 99% of the time the Montgomery code would be called) 022 */ 023 int 024 mp_exptmod (mp_int * G, mp_int * X, mp_int * P, mp_int * Y) 025 \{ 026 int dr; 027 028 /* modulus P must be positive */ 029 if (P->sign == MP_NEG) \{ 030 return MP_VAL; 031 \} 032 033 /* if exponent X is negative we have to recurse */ 034 if (X->sign == MP_NEG) \{ 035 mp_int tmpG, tmpX; 036 int err; 037 038 /* first compute 1/G mod P */ 039 if ((err = mp_init(&tmpG)) != MP_OKAY) \{ 040 return err; 041 \} 042 if ((err = mp_invmod(G, P, &tmpG)) != MP_OKAY) \{ 043 mp_clear(&tmpG); 044 return err; 045 \} 046 047 /* now get |X| */ 048 if ((err = mp_init(&tmpX)) != MP_OKAY) \{ 049 mp_clear(&tmpG); 050 return err; 051 \} 052 if ((err = mp_abs(X, &tmpX)) != MP_OKAY) \{ 053 mp_clear_multi(&tmpG, &tmpX, NULL); 054 return err; 055 \} 056 057 /* and now compute (1/G)**|X| instead of G**X [X < 0] */ 058 err = mp_exptmod(&tmpG, &tmpX, P, Y); 059 mp_clear_multi(&tmpG, &tmpX, NULL); 060 return err; 061 \} 062 063 dr = mp_dr_is_modulus(P); 064 if (dr == 0) \{ 065 dr = mp_reduce_is_2k(P) << 1; 066 \} 067 068 /* if the modulus is odd use the fast method */ 069 if ((mp_isodd (P) == 1 || dr != 0) && P->used > 4) \{ 070 return mp_exptmod_fast (G, X, P, Y, dr); 071 \} else \{ 072 return s_mp_exptmod (G, X, P, Y); 073 \} 074 \} 075 \end{alltt} \end{small} \subsection{Barrett Modular Exponentiation} \newpage\begin{figure}[!here] \begin{small} \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{s\_mp\_exptmod}. \\ \textbf{Input}. mp\_int $a$, $b$ and $c$ \\ \textbf{Output}. $y \equiv g^x \mbox{ (mod }p\mbox{)}$ \\ \hline \\ 1. $k \leftarrow lg(x)$ \\ 2. $winsize \leftarrow \left \lbrace \begin{array}{ll} 2 & \mbox{if }k \le 7 \\ 3 & \mbox{if }7 < k \le 36 \\ 4 & \mbox{if }36 < k \le 140 \\ 5 & \mbox{if }140 < k \le 450 \\ 6 & \mbox{if }450 < k \le 1303 \\ 7 & \mbox{if }1303 < k \le 3529 \\ 8 & \mbox{if }3529 < k \\ \end{array} \right .$ \\ 3. Initialize $2^{winsize}$ mp\_ints in an array named $M$ and one mp\_int named $\mu$ \\ 4. Calculate the $\mu$ required for Barrett Reduction (\textit{mp\_reduce\_setup}). \\ 5. $M_1 \leftarrow g \mbox{ (mod }p\mbox{)}$ \\ \\ Setup the table of small powers of $g$. First find $g^{2^{winsize}}$ and then all multiples of it. \\ 6. $k \leftarrow 2^{winsize - 1}$ \\ 7. $M_{k} \leftarrow M_1$ \\ 8. for $ix$ from 0 to $winsize - 2$ do \\ \hspace{3mm}8.1 $M_k \leftarrow \left ( M_k \right )^2$ \\ \hspace{3mm}8.2 $M_k \leftarrow M_k \mbox{ (mod }p\mbox{)}$ (\textit{mp\_reduce}) \\ 9. for $ix$ from $2^{winsize - 1} + 1$ to $2^{winsize} - 1$ do \\ \hspace{3mm}9.1 $M_{ix} \leftarrow M_{ix - 1} \cdot M_{1}$ \\ \hspace{3mm}9.2 $M_{ix} \leftarrow M_{ix} \mbox{ (mod }p\mbox{)}$ (\textit{mp\_reduce}) \\ 10. $res \leftarrow 1$ \\ \\ Start Sliding Window. \\ 11. $mode \leftarrow 0, bitcnt \leftarrow 1, buf \leftarrow 0, digidx \leftarrow x.used - 1, bitcpy \leftarrow 0, bitbuf \leftarrow 0$ \\ 12. Loop \\ \hspace{3mm}12.1 $bitcnt \leftarrow bitcnt - 1$ \\ \hspace{3mm}12.2 If $bitcnt = 0$ then do \\ \hspace{6mm}12.2.1 If $digidx = -1$ goto step 13. \\ \hspace{6mm}12.2.2 $buf \leftarrow x_{digidx}$ \\ \hspace{6mm}12.2.3 $digidx \leftarrow digidx - 1$ \\ \hspace{6mm}12.2.4 $bitcnt \leftarrow lg(\beta)$ \\ Continued on next page. \\ \hline \end{tabular} \end{center} \end{small} \caption{Algorithm s\_mp\_exptmod} \end{figure} \newpage\begin{figure}[!here] \begin{small} \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{s\_mp\_exptmod} (\textit{continued}). \\ \textbf{Input}. mp\_int $a$, $b$ and $c$ \\ \textbf{Output}. $y \equiv g^x \mbox{ (mod }p\mbox{)}$ \\ \hline \\ \hspace{3mm}12.3 $y \leftarrow (buf >> (lg(\beta) - 1))$ AND $1$ \\ \hspace{3mm}12.4 $buf \leftarrow buf << 1$ \\ \hspace{3mm}12.5 if $mode = 0$ and $y = 0$ then goto step 12. \\ \hspace{3mm}12.6 if $mode = 1$ and $y = 0$ then do \\ \hspace{6mm}12.6.1 $res \leftarrow res^2$ \\ \hspace{6mm}12.6.2 $res \leftarrow res \mbox{ (mod }p\mbox{)}$ \\ \hspace{6mm}12.6.3 Goto step 12. \\ \hspace{3mm}12.7 $bitcpy \leftarrow bitcpy + 1$ \\ \hspace{3mm}12.8 $bitbuf \leftarrow bitbuf + (y << (winsize - bitcpy))$ \\ \hspace{3mm}12.9 $mode \leftarrow 2$ \\ \hspace{3mm}12.10 If $bitcpy = winsize$ then do \\ \hspace{6mm}Window is full so perform the squarings and single multiplication. \\ \hspace{6mm}12.10.1 for $ix$ from $0$ to $winsize -1$ do \\ \hspace{9mm}12.10.1.1 $res \leftarrow res^2$ \\ \hspace{9mm}12.10.1.2 $res \leftarrow res \mbox{ (mod }p\mbox{)}$ \\ \hspace{6mm}12.10.2 $res \leftarrow res \cdot M_{bitbuf}$ \\ \hspace{6mm}12.10.3 $res \leftarrow res \mbox{ (mod }p\mbox{)}$ \\ \hspace{6mm}Reset the window. \\ \hspace{6mm}12.10.4 $bitcpy \leftarrow 0, bitbuf \leftarrow 0, mode \leftarrow 1$ \\ \\ No more windows left. Check for residual bits of exponent. \\ 13. If $mode = 2$ and $bitcpy > 0$ then do \\ \hspace{3mm}13.1 for $ix$ form $0$ to $bitcpy - 1$ do \\ \hspace{6mm}13.1.1 $res \leftarrow res^2$ \\ \hspace{6mm}13.1.2 $res \leftarrow res \mbox{ (mod }p\mbox{)}$ \\ \hspace{6mm}13.1.3 $bitbuf \leftarrow bitbuf << 1$ \\ \hspace{6mm}13.1.4 If $bitbuf$ AND $2^{winsize} \ne 0$ then do \\ \hspace{9mm}13.1.4.1 $res \leftarrow res \cdot M_{1}$ \\ \hspace{9mm}13.1.4.2 $res \leftarrow res \mbox{ (mod }p\mbox{)}$ \\ 14. $y \leftarrow res$ \\ 15. Clear $res$, $mu$ and the $M$ array. \\ 16. Return(\textit{MP\_OKAY}). \\ \hline \end{tabular} \end{center} \end{small} \caption{Algorithm s\_mp\_exptmod (continued)} \end{figure} \textbf{Algorithm s\_mp\_exptmod.} This algorithm computes the $x$'th power of $g$ modulo $p$ and stores the result in $y$. It takes advantage of the Barrett reduction algorithm to keep the product small throughout the algorithm. The first two steps determine the optimal window size based on the number of bits in the exponent. The larger the exponent the larger the window size becomes. After a window size $winsize$ has been chosen an array of $2^{winsize}$ mp\_int variables is allocated. This table will hold the values of $g^x \mbox{ (mod }p\mbox{)}$ for $2^{winsize - 1} \le x < 2^{winsize}$. After the table is allocated the first power of $g$ is found. Since $g \ge p$ is allowed it must be first reduced modulo $p$ to make the rest of the algorithm more efficient. The first element of the table at $2^{winsize - 1}$ is found by squaring $M_1$ successively $winsize - 2$ times. The rest of the table elements are found by multiplying the previous element by $M_1$ modulo $p$. Now that the table is available the sliding window may begin. The following list describes the functions of all the variables in the window. \begin{enumerate} \item The variable $mode$ dictates how the bits of the exponent are interpreted. \begin{enumerate} \item When $mode = 0$ the bits are ignored since no non-zero bit of the exponent has been seen yet. For example, if the exponent were simply $1$ then there would be $lg(\beta) - 1$ zero bits before the first non-zero bit. In this case bits are ignored until a non-zero bit is found. \item When $mode = 1$ a non-zero bit has been seen before and a new $winsize$-bit window has not been formed yet. In this mode leading $0$ bits are read and a single squaring is performed. If a non-zero bit is read a new window is created. \item When $mode = 2$ the algorithm is in the middle of forming a window and new bits are appended to the window from the most significant bit downards. \end{enumerate} \item The variable $bitcnt$ indicates how many bits are left in the current digit of the exponent left to be read. When it reaches zero a new digit is fetched from the exponent. \item The variable $buf$ holds the currently read digit of the exponent. \item The variable $digidx$ is an index into the exponents digits. It starts at the leading digit $x.used - 1$ and moves towards the trailing digit. \item The variable $bitcpy$ indicates how many bits are in the currently formed window. When it reaches $winsize$ the window is flushed and the appropriate operations performed. \item The variable $bitbuf$ holds the current bits of the window being formed. \end{enumerate} All of step 12 is the window processing loop. It will iterate while there are digits available form the exponent to read. The first step inside this loop is to extract a new digit if no more bits are available in the current digit. If there are no bits left a new digit is read and if there are no digits left than the loop terminates. After a digit is made available step 12.3 will extract the most significant bit of the current digit and move all other bits in the digit upwards. In effect the digit is read from most significant bit to least significant bit and since the digits are read from leading to trailing edges the entire exponent is read from most significant bit to least significant bit. At step 12.5 if the $mode$ and currently extracted bit $y$ are both zero the bit is ignored and the next bit is read. This prevents the algorithm from having todo trivial squaring and reduction operations before the first non-zero bit is read. Step 12.6 and 12.7-10 handle the two cases of $mode = 1$ and $mode = 2$ respectively. \begin{center} \begin{figure}[here] \includegraphics{pics/expt_state.ps} \caption{Sliding Window State Diagram} \end{figure} \end{center} By step 13 there are no more digits left in the exponent. However, there may be partial bits in the window left. If $mode = 2$ then a Left-to-Right algorithm is used to process the remaining few bits. \index{bn\_s\_mp\_exptmod.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_s\_mp\_exptmod.c \vspace{-3mm} \begin{alltt} 016 017 int 018 s_mp_exptmod (mp_int * G, mp_int * X, mp_int * P, mp_int * Y) 019 \{ 020 mp_int M[256], res, mu; 021 mp_digit buf; 022 int err, bitbuf, bitcpy, bitcnt, mode, digidx, x, y, winsize; 023 024 /* find window size */ 025 x = mp_count_bits (X); 026 if (x <= 7) \{ 027 winsize = 2; 028 \} else if (x <= 36) \{ 029 winsize = 3; 030 \} else if (x <= 140) \{ 031 winsize = 4; 032 \} else if (x <= 450) \{ 033 winsize = 5; 034 \} else if (x <= 1303) \{ 035 winsize = 6; 036 \} else if (x <= 3529) \{ 037 winsize = 7; 038 \} else \{ 039 winsize = 8; 040 \} 041 042 #ifdef MP_LOW_MEM 043 if (winsize > 5) \{ 044 winsize = 5; 045 \} 046 #endif 047 048 /* init M array */ 049 for (x = 0; x < (1 << winsize); x++) \{ 050 if ((err = mp_init_size (&M[x], 1)) != MP_OKAY) \{ 051 for (y = 0; y < x; y++) \{ 052 mp_clear (&M[y]); 053 \} 054 return err; 055 \} 056 \} 057 058 /* create mu, used for Barrett reduction */ 059 if ((err = mp_init (&mu)) != MP_OKAY) \{ 060 goto __M; 061 \} 062 if ((err = mp_reduce_setup (&mu, P)) != MP_OKAY) \{ 063 goto __MU; 064 \} 065 066 /* create M table 067 * 068 * The M table contains powers of the input base, e.g. M[x] = G**x mod P 069 * 070 * The first half of the table is not computed though accept for M[0] and M[1] 071 */ 072 if ((err = mp_mod (G, P, &M[1])) != MP_OKAY) \{ 073 goto __MU; 074 \} 075 076 /* compute the value at M[1<<(winsize-1)] by squaring M[1] (winsize-1) tim es */ 077 if ((err = mp_copy (&M[1], &M[1 << (winsize - 1)])) != MP_OKAY) \{ 078 goto __MU; 079 \} 080 081 for (x = 0; x < (winsize - 1); x++) \{ 082 if ((err = mp_sqr (&M[1 << (winsize - 1)], &M[1 << (winsize - 1)])) != M P_OKAY) \{ 083 goto __MU; 084 \} 085 if ((err = mp_reduce (&M[1 << (winsize - 1)], P, &mu)) != MP_OKAY) \{ 086 goto __MU; 087 \} 088 \} 089 090 /* create upper table */ 091 for (x = (1 << (winsize - 1)) + 1; x < (1 << winsize); x++) \{ 092 if ((err = mp_mul (&M[x - 1], &M[1], &M[x])) != MP_OKAY) \{ 093 goto __MU; 094 \} 095 if ((err = mp_reduce (&M[x], P, &mu)) != MP_OKAY) \{ 096 goto __MU; 097 \} 098 \} 099 100 /* setup result */ 101 if ((err = mp_init (&res)) != MP_OKAY) \{ 102 goto __MU; 103 \} 104 mp_set (&res, 1); 105 106 /* set initial mode and bit cnt */ 107 mode = 0; 108 bitcnt = 1; 109 buf = 0; 110 digidx = X->used - 1; 111 bitcpy = bitbuf = 0; 112 113 for (;;) \{ 114 /* grab next digit as required */ 115 if (--bitcnt == 0) \{ 116 if (digidx == -1) \{ 117 break; 118 \} 119 buf = X->dp[digidx--]; 120 bitcnt = (int) DIGIT_BIT; 121 \} 122 123 /* grab the next msb from the exponent */ 124 y = (buf >> (mp_digit)(DIGIT_BIT - 1)) & 1; 125 buf <<= (mp_digit)1; 126 127 /* if the bit is zero and mode == 0 then we ignore it 128 * These represent the leading zero bits before the first 1 bit 129 * in the exponent. Technically this opt is not required but it 130 * does lower the # of trivial squaring/reductions used 131 */ 132 if (mode == 0 && y == 0) 133 continue; 134 135 /* if the bit is zero and mode == 1 then we square */ 136 if (mode == 1 && y == 0) \{ 137 if ((err = mp_sqr (&res, &res)) != MP_OKAY) \{ 138 goto __RES; 139 \} 140 if ((err = mp_reduce (&res, P, &mu)) != MP_OKAY) \{ 141 goto __RES; 142 \} 143 continue; 144 \} 145 146 /* else we add it to the window */ 147 bitbuf |= (y << (winsize - ++bitcpy)); 148 mode = 2; 149 150 if (bitcpy == winsize) \{ 151 /* ok window is filled so square as required and multiply */ 152 /* square first */ 153 for (x = 0; x < winsize; x++) \{ 154 if ((err = mp_sqr (&res, &res)) != MP_OKAY) \{ 155 goto __RES; 156 \} 157 if ((err = mp_reduce (&res, P, &mu)) != MP_OKAY) \{ 158 goto __RES; 159 \} 160 \} 161 162 /* then multiply */ 163 if ((err = mp_mul (&res, &M[bitbuf], &res)) != MP_OKAY) \{ 164 goto __MU; 165 \} 166 if ((err = mp_reduce (&res, P, &mu)) != MP_OKAY) \{ 167 goto __MU; 168 \} 169 170 /* empty window and reset */ 171 bitcpy = bitbuf = 0; 172 mode = 1; 173 \} 174 \} 175 176 /* if bits remain then square/multiply */ 177 if (mode == 2 && bitcpy > 0) \{ 178 /* square then multiply if the bit is set */ 179 for (x = 0; x < bitcpy; x++) \{ 180 if ((err = mp_sqr (&res, &res)) != MP_OKAY) \{ 181 goto __RES; 182 \} 183 if ((err = mp_reduce (&res, P, &mu)) != MP_OKAY) \{ 184 goto __RES; 185 \} 186 187 bitbuf <<= 1; 188 if ((bitbuf & (1 << winsize)) != 0) \{ 189 /* then multiply */ 190 if ((err = mp_mul (&res, &M[1], &res)) != MP_OKAY) \{ 191 goto __RES; 192 \} 193 if ((err = mp_reduce (&res, P, &mu)) != MP_OKAY) \{ 194 goto __RES; 195 \} 196 \} 197 \} 198 \} 199 200 mp_exch (&res, Y); 201 err = MP_OKAY; 202 __RES:mp_clear (&res); 203 __MU:mp_clear (&mu); 204 __M: 205 for (x = 0; x < (1 << winsize); x++) \{ 206 mp_clear (&M[x]); 207 \} 208 return err; 209 \} \end{alltt} \end{small} \section{Quick Power of Two} Calculating $b = 2^a$ can be performed much quicker than with any of the previous algorithms. Recall that a logical shift left $m << k$ is equivalent to $m \cdot 2^k$. By this logic when $m = 1$ a quick power of two can be achieved. \newpage\begin{figure}[!here] \begin{small} \begin{center} \begin{tabular}{l} \hline Algorithm \textbf{mp\_2expt}. \\ \textbf{Input}. integer $b$ \\ \textbf{Output}. $a \leftarrow 2^b$ \\ \hline \\ 1. $a \leftarrow 0$ \\ 2. If $a.alloc < \lfloor b / lg(\beta) \rfloor + 1$ then grow $a$ appropriately. \\ 3. $a.used \leftarrow \lfloor b / lg(\beta) \rfloor + 1$ \\ 4. $a_{\lfloor b / lg(\beta) \rfloor} \leftarrow 1 << (b \mbox{ mod } lg(\beta))$ \\ 5. Return(\textit{MP\_OKAY}). \\ \hline \end{tabular} \end{center} \end{small} \caption{Algorithm mp\_2expt} \end{figure} \textbf{Algorithm mp\_2expt.} \index{bn\_mp\_2expt.c} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: bn\_mp\_2expt.c \vspace{-3mm} \begin{alltt} 016 017 /* computes a = 2**b 018 * 019 * Simple algorithm which zeroes the int, grows it then just sets one bit 020 * as required. 021 */ 022 int 023 mp_2expt (mp_int * a, int b) 024 \{ 025 int res; 026 027 mp_zero (a); 028 if ((res = mp_grow (a, b / DIGIT_BIT + 1)) != MP_OKAY) \{ 029 return res; 030 \} 031 a->used = b / DIGIT_BIT + 1; 032 a->dp[b / DIGIT_BIT] = 1 << (b % DIGIT_BIT); 033 034 return MP_OKAY; 035 \} \end{alltt} \end{small} \chapter{Higher Level Algorithms} \section{Integer Division with Remainder} \section{Single Digit Helpers} \subsection{Single Digit Addition} \subsection{Single Digit Subtraction} \subsection{Single Digit Multiplication} \subsection{Single Digit Division} \subsection{Single Digit Modulo} \subsection{Single Digit Root Extraction} \section{Random Number Generation} \section{Formatted Output} \subsection{Getting The Output Size} \subsection{Generating Radix-n Output} \subsection{Reading Radix-n Input} \section{Unformatted Output} \subsection{Getting The Output Size} \subsection{Generating Output} \subsection{Reading Input} \chapter{Number Theoretic Algorithms} \section{Greatest Common Divisor} \section{Least Common Multiple} \section{Jacobi Symbol Computation} \section{Modular Inverse} \subsection{General Case} \subsection{Odd Moduli} \section{Primality Tests} \subsection{Trial Division} \subsection{The Fermat Test} \subsection{The Miller-Rabin Test} \subsection{Primality Test in a Bottle} \subsection{The Next Prime} \section{Root Extraction} \backmatter \appendix \begin{thebibliography}{ABCDEF} \bibitem[1]{TAOCPV2} Donald Knuth, \textit{The Art of Computer Programming}, Third Edition, Volume Two, Seminumerical Algorithms, Addison-Wesley, 1998 \bibitem[2]{HAC} A. Menezes, P. van Oorschot, S. Vanstone, \textit{Handbook of Applied Cryptography}, CRC Press, 1996 \bibitem[3]{ROSE} Michael Rosing, \textit{Implementing Elliptic Curve Cryptography}, Manning Publications, 1999 \bibitem[4]{COMBA} Paul G. Comba, \textit{Exponentiation Cryptosystems on the IBM PC}. IBM Systems Journal 29(4): 526-538 (1990) \bibitem[5]{KARA} A. Karatsuba, Doklay Akad. Nauk SSSR 145 (1962), pp.293-294 \bibitem[6]{KARAP} Andre Weimerskirch and Christof Paar, \textit{Generalizations of the Karatsuba Algorithm for Polynomial Multiplication}, Submitted to Design, Codes and Cryptography, March 2002 \bibitem[7]{BARRETT} Paul Barrett, \textit{Implementing the Rivest Shamir and Adleman Public Key Encryption Algorithm on a Standard Digital Signal Processor}, Advances in Cryptology, Crypto '86, Springer-Verlag. \bibitem[8]{MONT} P.L.Montgomery. \textit{Modular multiplication without trial division}. Mathematics of Computation, 44(170):519-521, April 1985. \bibitem[9]{DRMET} Chae Hoon Lim and Pil Joong Lee, \textit{Generating Efficient Primes for Discrete Log Cryptosystems}, POSTECH Information Research Laboratories \bibitem[10]{MMB} J. Daemen and R. Govaerts and J. Vandewalle, \textit{Block ciphers based on Modular Arithmetic}, State and {P}rogress in the {R}esearch of {C}ryptography, 1993, pp. 80-89 \end{thebibliography} \input{tommath.ind} \chapter{Appendix} \subsection*{Appendix A -- Source Listing of tommath.h} The following is the source listing of the header file ``tommath.h'' for the LibTomMath project. It contains many of the definitions used throughout the code such as \textbf{mp\_int}, \textbf{MP\_PREC} and so on. The header is presented here for completeness. \index{tommath.h} \vspace{+3mm}\begin{small} \hspace{-5.1mm}{\bf File}: tommath.h \vspace{-3mm} \begin{alltt} 001 /* LibTomMath, multiple-precision integer library -- Tom St Denis 002 * 003 * LibTomMath is library that provides for multiple-precision 004 * integer arithmetic as well as number theoretic functionality. 005 * 006 * The library is designed directly after the MPI library by 007 * Michael Fromberger but has been written from scratch with 008 * additional optimizations in place. 009 * 010 * The library is free for all purposes without any express 011 * guarantee it works. 012 * 013 * Tom St Denis, tomstdenis@iahu.ca, http://math.libtomcrypt.org 014 */ 015 #ifndef BN_H_ 016 #define BN_H_ 017 018 #include 019 #include 020 #include 021 #include 022 #include 023 024 #undef MIN 025 #define MIN(x,y) ((x)<(y)?(x):(y)) 026 #undef MAX 027 #define MAX(x,y) ((x)>(y)?(x):(y)) 028 029 #ifdef __cplusplus 030 extern "C" \{ 031 032 /* C++ compilers don't like assigning void * to mp_digit * */ 033 #define OPT_CAST (mp_digit *) 034 035 #else 036 037 /* C on the other hand doesn't care */ 038 #define OPT_CAST 039 040 #endif 041 042 /* some default configurations. 043 * 044 * A "mp_digit" must be able to hold DIGIT_BIT + 1 bits 045 * A "mp_word" must be able to hold 2*DIGIT_BIT + 1 bits 046 * 047 * At the very least a mp_digit must be able to hold 7 bits 048 * [any size beyond that is ok provided it doesn't overflow the data type] 049 */ 050 #ifdef MP_8BIT 051 typedef unsigned char mp_digit; 052 typedef unsigned short mp_word; 053 #elif defined(MP_16BIT) 054 typedef unsigned short mp_digit; 055 typedef unsigned long mp_word; 056 #elif defined(MP_64BIT) 057 /* for GCC only on supported platforms */ 058 #ifndef CRYPT 059 typedef unsigned long long ulong64; 060 typedef signed long long long64; 061 #endif 062 063 typedef ulong64 mp_digit; 064 typedef unsigned long mp_word __attribute__ ((mode(TI))); 065 066 #define DIGIT_BIT 60 067 #else 068 /* this is the default case, 28-bit digits */ 069 070 /* this is to make porting into LibTomCrypt easier :-) */ 071 #ifndef CRYPT 072 #if defined(_MSC_VER) || defined(__BORLANDC__) 073 typedef unsigned __int64 ulong64; 074 typedef signed __int64 long64; 075 #else 076 typedef unsigned long long ulong64; 077 typedef signed long long long64; 078 #endif 079 #endif 080 081 typedef unsigned long mp_digit; 082 typedef ulong64 mp_word; 083 084 #ifdef MP_31BIT 085 #define DIGIT_BIT 31 086 #else 087 #define DIGIT_BIT 28 088 #endif 089 #endif 090 091 /* otherwise the bits per digit is calculated automatically from the size of a mp_digit */ 092 #ifndef DIGIT_BIT 093 #define DIGIT_BIT ((CHAR_BIT * sizeof(mp_digit) - 1)) /* bits per di git */ 094 #endif 095 096 097 #define MP_DIGIT_BIT DIGIT_BIT 098 #define MP_MASK ((((mp_digit)1)<<((mp_digit)DIGIT_BIT))-((mp_digit) 1)) 099 #define MP_DIGIT_MAX MP_MASK 100 101 /* equalities */ 102 #define MP_LT -1 /* less than */ 103 #define MP_EQ 0 /* equal to */ 104 #define MP_GT 1 /* greater than */ 105 106 #define MP_ZPOS 0 /* positive integer */ 107 #define MP_NEG 1 /* negative */ 108 109 #define MP_OKAY 0 /* ok result */ 110 #define MP_MEM -2 /* out of mem */ 111 #define MP_VAL -3 /* invalid input */ 112 #define MP_RANGE MP_VAL 113 114 typedef int mp_err; 115 116 /* you'll have to tune these... */ 117 extern int KARATSUBA_MUL_CUTOFF, 118 KARATSUBA_SQR_CUTOFF, 119 TOOM_MUL_CUTOFF, 120 TOOM_SQR_CUTOFF; 121 122 /* various build options */ 123 #define MP_PREC 64 /* default digits of precision (must be power of two) */ 124 125 /* define this to use lower memory usage routines (exptmods mostly) */ 126 /* #define MP_LOW_MEM */ 127 128 /* size of comba arrays, should be at least 2 * 2**(BITS_PER_WORD - BITS_PER _DIGIT*2) */ 129 #define MP_WARRAY (1 << (sizeof(mp_word) * CHAR_BIT - 2 * DIGI T_BIT + 1)) 130 131 typedef struct \{ 132 int used, alloc, sign; 133 mp_digit *dp; 134 \} mp_int; 135 136 #define USED(m) ((m)->used) 137 #define DIGIT(m,k) ((m)->dp[k]) 138 #define SIGN(m) ((m)->sign) 139 140 /* ---> init and deinit bignum functions <--- */ 141 142 /* init a bignum */ 143 int mp_init(mp_int *a); 144 145 /* free a bignum */ 146 void mp_clear(mp_int *a); 147 148 /* init a null terminated series of arguments */ 149 int mp_init_multi(mp_int *mp, ...); 150 151 /* clear a null terminated series of arguments */ 152 void mp_clear_multi(mp_int *mp, ...); 153 154 /* exchange two ints */ 155 void mp_exch(mp_int *a, mp_int *b); 156 157 /* shrink ram required for a bignum */ 158 int mp_shrink(mp_int *a); 159 160 /* grow an int to a given size */ 161 int mp_grow(mp_int *a, int size); 162 163 /* init to a given number of digits */ 164 int mp_init_size(mp_int *a, int size); 165 166 /* ---> Basic Manipulations <--- */ 167 168 #define mp_iszero(a) (((a)->used == 0) ? 1 : 0) 169 #define mp_iseven(a) (((a)->used == 0 || (((a)->dp[0] & 1) == 0)) ? 1 : 0) 170 #define mp_isodd(a) (((a)->used > 0 && (((a)->dp[0] & 1) == 1)) ? 1 : 0) 171 172 /* set to zero */ 173 void mp_zero(mp_int *a); 174 175 /* set to a digit */ 176 void mp_set(mp_int *a, mp_digit b); 177 178 /* set a 32-bit const */ 179 int mp_set_int(mp_int *a, unsigned int b); 180 181 /* copy, b = a */ 182 int mp_copy(mp_int *a, mp_int *b); 183 184 /* inits and copies, a = b */ 185 int mp_init_copy(mp_int *a, mp_int *b); 186 187 /* trim unused digits */ 188 void mp_clamp(mp_int *a); 189 190 /* ---> digit manipulation <--- */ 191 192 /* right shift by "b" digits */ 193 void mp_rshd(mp_int *a, int b); 194 195 /* left shift by "b" digits */ 196 int mp_lshd(mp_int *a, int b); 197 198 /* c = a / 2**b */ 199 int mp_div_2d(mp_int *a, int b, mp_int *c, mp_int *d); 200 201 /* b = a/2 */ 202 int mp_div_2(mp_int *a, mp_int *b); 203 204 /* c = a * 2**b */ 205 int mp_mul_2d(mp_int *a, int b, mp_int *c); 206 207 /* b = a*2 */ 208 int mp_mul_2(mp_int *a, mp_int *b); 209 210 /* c = a mod 2**d */ 211 int mp_mod_2d(mp_int *a, int b, mp_int *c); 212 213 /* computes a = 2**b */ 214 int mp_2expt(mp_int *a, int b); 215 216 /* makes a pseudo-random int of a given size */ 217 int mp_rand(mp_int *a, int digits); 218 219 /* ---> binary operations <--- */ 220 /* c = a XOR b */ 221 int mp_xor(mp_int *a, mp_int *b, mp_int *c); 222 223 /* c = a OR b */ 224 int mp_or(mp_int *a, mp_int *b, mp_int *c); 225 226 /* c = a AND b */ 227 int mp_and(mp_int *a, mp_int *b, mp_int *c); 228 229 /* ---> Basic arithmetic <--- */ 230 231 /* b = -a */ 232 int mp_neg(mp_int *a, mp_int *b); 233 234 /* b = |a| */ 235 int mp_abs(mp_int *a, mp_int *b); 236 237 /* compare a to b */ 238 int mp_cmp(mp_int *a, mp_int *b); 239 240 /* compare |a| to |b| */ 241 int mp_cmp_mag(mp_int *a, mp_int *b); 242 243 /* c = a + b */ 244 int mp_add(mp_int *a, mp_int *b, mp_int *c); 245 246 /* c = a - b */ 247 int mp_sub(mp_int *a, mp_int *b, mp_int *c); 248 249 /* c = a * b */ 250 int mp_mul(mp_int *a, mp_int *b, mp_int *c); 251 252 /* b = a*a */ 253 int mp_sqr(mp_int *a, mp_int *b); 254 255 /* a/b => cb + d == a */ 256 int mp_div(mp_int *a, mp_int *b, mp_int *c, mp_int *d); 257 258 /* c = a mod b, 0 <= c < b */ 259 int mp_mod(mp_int *a, mp_int *b, mp_int *c); 260 261 /* ---> single digit functions <--- */ 262 263 /* compare against a single digit */ 264 int mp_cmp_d(mp_int *a, mp_digit b); 265 266 /* c = a + b */ 267 int mp_add_d(mp_int *a, mp_digit b, mp_int *c); 268 269 /* c = a - b */ 270 int mp_sub_d(mp_int *a, mp_digit b, mp_int *c); 271 272 /* c = a * b */ 273 int mp_mul_d(mp_int *a, mp_digit b, mp_int *c); 274 275 /* a/b => cb + d == a */ 276 int mp_div_d(mp_int *a, mp_digit b, mp_int *c, mp_digit *d); 277 278 /* a/3 => 3c + d == a */ 279 int mp_div_3(mp_int *a, mp_int *c, mp_digit *d); 280 281 /* c = a**b */ 282 int mp_expt_d(mp_int *a, mp_digit b, mp_int *c); 283 284 /* c = a mod b, 0 <= c < b */ 285 int mp_mod_d(mp_int *a, mp_digit b, mp_digit *c); 286 287 /* ---> number theory <--- */ 288 289 /* d = a + b (mod c) */ 290 int mp_addmod(mp_int *a, mp_int *b, mp_int *c, mp_int *d); 291 292 /* d = a - b (mod c) */ 293 int mp_submod(mp_int *a, mp_int *b, mp_int *c, mp_int *d); 294 295 /* d = a * b (mod c) */ 296 int mp_mulmod(mp_int *a, mp_int *b, mp_int *c, mp_int *d); 297 298 /* c = a * a (mod b) */ 299 int mp_sqrmod(mp_int *a, mp_int *b, mp_int *c); 300 301 /* c = 1/a (mod b) */ 302 int mp_invmod(mp_int *a, mp_int *b, mp_int *c); 303 304 /* c = (a, b) */ 305 int mp_gcd(mp_int *a, mp_int *b, mp_int *c); 306 307 /* c = [a, b] or (a*b)/(a, b) */ 308 int mp_lcm(mp_int *a, mp_int *b, mp_int *c); 309 310 /* finds one of the b'th root of a, such that |c|**b <= |a| 311 * 312 * returns error if a < 0 and b is even 313 */ 314 int mp_n_root(mp_int *a, mp_digit b, mp_int *c); 315 316 /* shortcut for square root */ 317 #define mp_sqrt(a, b) mp_n_root(a, 2, b) 318 319 /* computes the jacobi c = (a | n) (or Legendre if b is prime) */ 320 int mp_jacobi(mp_int *a, mp_int *n, int *c); 321 322 /* used to setup the Barrett reduction for a given modulus b */ 323 int mp_reduce_setup(mp_int *a, mp_int *b); 324 325 /* Barrett Reduction, computes a (mod b) with a precomputed value c 326 * 327 * Assumes that 0 < a <= b*b, note if 0 > a > -(b*b) then you can merely 328 * compute the reduction as -1 * mp_reduce(mp_abs(a)) [pseudo code]. 329 */ 330 int mp_reduce(mp_int *a, mp_int *b, mp_int *c); 331 332 /* setups the montgomery reduction */ 333 int mp_montgomery_setup(mp_int *a, mp_digit *mp); 334 335 /* computes a = B**n mod b without division or multiplication useful for 336 * normalizing numbers in a Montgomery system. 337 */ 338 int mp_montgomery_calc_normalization(mp_int *a, mp_int *b); 339 340 /* computes x/R == x (mod N) via Montgomery Reduction */ 341 int mp_montgomery_reduce(mp_int *a, mp_int *m, mp_digit mp); 342 343 /* returns 1 if a is a valid DR modulus */ 344 int mp_dr_is_modulus(mp_int *a); 345 346 /* sets the value of "d" required for mp_dr_reduce */ 347 void mp_dr_setup(mp_int *a, mp_digit *d); 348 349 /* reduces a modulo b using the Diminished Radix method */ 350 int mp_dr_reduce(mp_int *a, mp_int *b, mp_digit mp); 351 352 /* returns true if a can be reduced with mp_reduce_2k */ 353 int mp_reduce_is_2k(mp_int *a); 354 355 /* determines k value for 2k reduction */ 356 int mp_reduce_2k_setup(mp_int *a, mp_digit *d); 357 358 /* reduces a modulo b where b is of the form 2**p - k [0 <= a] */ 359 int mp_reduce_2k(mp_int *a, mp_int *n, mp_digit k); 360 361 /* d = a**b (mod c) */ 362 int mp_exptmod(mp_int *a, mp_int *b, mp_int *c, mp_int *d); 363 364 /* ---> Primes <--- */ 365 366 /* number of primes */ 367 #ifdef MP_8BIT 368 #define PRIME_SIZE 31 369 #else 370 #define PRIME_SIZE 256 371 #endif 372 373 /* table of first PRIME_SIZE primes */ 374 extern const mp_digit __prime_tab[]; 375 376 /* result=1 if a is divisible by one of the first PRIME_SIZE primes */ 377 int mp_prime_is_divisible(mp_int *a, int *result); 378 379 /* performs one Fermat test of "a" using base "b". 380 * Sets result to 0 if composite or 1 if probable prime 381 */ 382 int mp_prime_fermat(mp_int *a, mp_int *b, int *result); 383 384 /* performs one Miller-Rabin test of "a" using base "b". 385 * Sets result to 0 if composite or 1 if probable prime 386 */ 387 int mp_prime_miller_rabin(mp_int *a, mp_int *b, int *result); 388 389 /* performs t rounds of Miller-Rabin on "a" using the first 390 * t prime bases. Also performs an initial sieve of trial 391 * division. Determines if "a" is prime with probability 392 * of error no more than (1/4)**t. 393 * 394 * Sets result to 1 if probably prime, 0 otherwise 395 */ 396 int mp_prime_is_prime(mp_int *a, int t, int *result); 397 398 /* finds the next prime after the number "a" using "t" trials 399 * of Miller-Rabin. 400 */ 401 int mp_prime_next_prime(mp_int *a, int t); 402 403 404 /* ---> radix conversion <--- */ 405 int mp_count_bits(mp_int *a); 406 407 int mp_unsigned_bin_size(mp_int *a); 408 int mp_read_unsigned_bin(mp_int *a, unsigned char *b, int c); 409 int mp_to_unsigned_bin(mp_int *a, unsigned char *b); 410 411 int mp_signed_bin_size(mp_int *a); 412 int mp_read_signed_bin(mp_int *a, unsigned char *b, int c); 413 int mp_to_signed_bin(mp_int *a, unsigned char *b); 414 415 int mp_read_radix(mp_int *a, char *str, int radix); 416 int mp_toradix(mp_int *a, char *str, int radix); 417 int mp_radix_size(mp_int *a, int radix); 418 419 int mp_fread(mp_int *a, int radix, FILE *stream); 420 int mp_fwrite(mp_int *a, int radix, FILE *stream); 421 422 #define mp_read_raw(mp, str, len) mp_read_signed_bin((mp), (str), (len)) 423 #define mp_raw_size(mp) mp_signed_bin_size(mp) 424 #define mp_toraw(mp, str) mp_to_signed_bin((mp), (str)) 425 #define mp_read_mag(mp, str, len) mp_read_unsigned_bin((mp), (str), (len)) 426 #define mp_mag_size(mp) mp_unsigned_bin_size(mp) 427 #define mp_tomag(mp, str) mp_to_unsigned_bin((mp), (str)) 428 429 #define mp_tobinary(M, S) mp_toradix((M), (S), 2) 430 #define mp_tooctal(M, S) mp_toradix((M), (S), 8) 431 #define mp_todecimal(M, S) mp_toradix((M), (S), 10) 432 #define mp_tohex(M, S) mp_toradix((M), (S), 16) 433 434 /* lowlevel functions, do not call! */ 435 int s_mp_add(mp_int *a, mp_int *b, mp_int *c); 436 int s_mp_sub(mp_int *a, mp_int *b, mp_int *c); 437 #define s_mp_mul(a, b, c) s_mp_mul_digs(a, b, c, (a)->used + (b)->used + 1) 438 int fast_s_mp_mul_digs(mp_int *a, mp_int *b, mp_int *c, int digs); 439 int s_mp_mul_digs(mp_int *a, mp_int *b, mp_int *c, int digs); 440 int fast_s_mp_mul_high_digs(mp_int *a, mp_int *b, mp_int *c, int digs); 441 int s_mp_mul_high_digs(mp_int *a, mp_int *b, mp_int *c, int digs); 442 int fast_s_mp_sqr(mp_int *a, mp_int *b); 443 int s_mp_sqr(mp_int *a, mp_int *b); 444 int mp_karatsuba_mul(mp_int *a, mp_int *b, mp_int *c); 445 int mp_toom_mul(mp_int *a, mp_int *b, mp_int *c); 446 int mp_karatsuba_sqr(mp_int *a, mp_int *b); 447 int mp_toom_sqr(mp_int *a, mp_int *b); 448 int fast_mp_invmod(mp_int *a, mp_int *b, mp_int *c); 449 int fast_mp_montgomery_reduce(mp_int *a, mp_int *m, mp_digit mp); 450 int mp_exptmod_fast(mp_int *G, mp_int *X, mp_int *P, mp_int *Y, int mode); 451 int s_mp_exptmod (mp_int * G, mp_int * X, mp_int * P, mp_int * Y); 452 void bn_reverse(unsigned char *s, int len); 453 454 #ifdef __cplusplus 455 \} 456 #endif 457 458 #endif 459 \end{alltt} \end{small} \end{document}