Modular multiplicative inverse

Last updated

In mathematics, particularly in the area of arithmetic, a modular multiplicative inverse of an integer a is an integer x such that the product ax is congruent to 1 with respect to the modulus m. [1] In the standard notation of modular arithmetic this congruence is written as

Contents

which is the shorthand way of writing the statement that m divides (evenly) the quantity ax − 1, or, put another way, the remainder after dividing ax by the integer m is 1. If a does have an inverse modulo m, then there are an infinite number of solutions of this congruence, which form a congruence class with respect to this modulus. Furthermore, any integer that is congruent to a (i.e., in a's congruence class) has any element of x's congruence class as a modular multiplicative inverse. Using the notation of to indicate the congruence class containing w, this can be expressed by saying that the modulo multiplicative inverse of the congruence class is the congruence class such that:

where the symbol denotes the multiplication of equivalence classes modulo m. [2] Written in this way, the analogy with the usual concept of a multiplicative inverse in the set of rational or real numbers is clearly represented, replacing the numbers by congruence classes and altering the binary operation appropriately.

As with the analogous operation on the real numbers, a fundamental use of this operation is in solving, when possible, linear congruences of the form

Finding modular multiplicative inverses also has practical applications in the field of cryptography, e.g. public-key cryptography and the RSA algorithm. [3] [4] [5] A benefit for the computer implementation of these applications is that there exists a very fast algorithm (the extended Euclidean algorithm) that can be used for the calculation of modular multiplicative inverses.

Modular arithmetic

For a given positive integer m, two integers, a and b, are said to be congruent modulo m if m divides their difference. This binary relation is denoted by,

This is an equivalence relation on the set of integers, , and the equivalence classes are called congruence classes modulo m or residue classes modulo m. Let denote the congruence class containing the integer a, [6] then

A linear congruence is a modular congruence of the form

Unlike linear equations over the reals, linear congruences may have zero, one or several solutions. If x is a solution of a linear congruence then every element in is also a solution, so, when speaking of the number of solutions of a linear congruence we are referring to the number of different congruence classes that contain solutions.

If d is the greatest common divisor of a and m then the linear congruence axb (mod m) has solutions if and only if d divides b. If d divides b, then there are exactly d solutions. [7]

A modular multiplicative inverse of an integer a with respect to the modulus m is a solution of the linear congruence

The previous result says that a solution exists if and only if gcd(a, m) = 1, that is, a and m must be relatively prime (i.e. coprime). Furthermore, when this condition holds, there is exactly one solution, i.e., when it exists, a modular multiplicative inverse is unique: [8] If b and b' are both modular multiplicative inverses of a respect to the modulus m, then

therefore

If a ≡ 0 (mod m), then gcd(a, m) = a, and a won't even have a modular multiplicative inverse. Therefore, b ≡ b' (mod m).

When ax ≡ 1 (mod m) has a solution it is often denoted in this way −

but this can be considered an abuse of notation since it could be misinterpreted as the reciprocal of (which, contrary to the modular multiplicative inverse, is not an integer except when a is 1 or -1). The notation would be proper if a is interpreted as a token standing for the congruence class , as the multiplicative inverse of a congruence class is a congruence class with the multiplication defined in the next section.

Integers modulo m

The congruence relation, modulo m, partitions the set of integers into m congruence classes. Operations of addition and multiplication can be defined on these m objects in the following way: To either add or multiply two congruence classes, first pick a representative (in any way) from each class, then perform the usual operation for integers on the two representatives and finally take the congruence class that the result of the integer operation lies in as the result of the operation on the congruence classes. In symbols, with and representing the operations on congruence classes, these definitions are

and

These operations are well-defined, meaning that the end result does not depend on the choices of representatives that were made to obtain the result.

The m congruence classes with these two defined operations form a ring, called the ring of integers modulo m. There are several notations used for these algebraic objects, most often or , but several elementary texts and application areas use a simplified notation when confusion with other algebraic objects is unlikely.

The congruence classes of the integers modulo m were traditionally known as residue classes modulo m, reflecting the fact that all the elements of a congruence class have the same remainder (i.e., "residue") upon being divided by m. Any set of m integers selected so that each comes from a different congruence class modulo m is called a complete system of residues modulo m. [9] The division algorithm shows that the set of integers, {0, 1, 2, ..., m − 1} form a complete system of residues modulo m, known as the least residue system modulo m. In working with arithmetic problems it is sometimes more convenient to work with a complete system of residues and use the language of congruences while at other times the point of view of the congruence classes of the ring is more useful. [10]

Multiplicative group of integers modulo m

Not every element of a complete residue system modulo m has a modular multiplicative inverse, for instance, zero never does. After removing the elements of a complete residue system that are not relatively prime to m, what is left is called a reduced residue system , all of whose elements have modular multiplicative inverses. The number of elements in a reduced residue system is , where is the Euler totient function, i.e., the number of positive integers less than m that are relatively prime to m.

In a general ring with unity not every element has a multiplicative inverse and those that do are called units. As the product of two units is a unit, the units of a ring form a group, the group of units of the ring and often denoted by R× if R is the name of the ring. The group of units of the ring of integers modulo m is called the multiplicative group of integers modulo m, and it is isomorphic to a reduced residue system. In particular, it has order (size), .

In the case that m is a prime, say p, then and all the non-zero elements of have multiplicative inverses, thus is a finite field. In this case, the multiplicative group of integers modulo p form a cyclic group of order p − 1.

Example

For any integer , it's always the case that is the modular multiplicative inverse of with respect to the modulus , since . Examples are , , and so on.

The following example uses the modulus 10: Two integers are congruent mod 10 if and only if their difference is divisible by 10, for instance

since 10 divides 32 − 2 = 30, and
since 10 divides 111 − 1 = 110.

Some of the ten congruence classes with respect to this modulus are:

and

The linear congruence 4x ≡ 5 (mod 10) has no solutions since the integers that are congruent to 5 (i.e., those in ) are all odd while 4x is always even. However, the linear congruence 4x ≡ 6 (mod 10) has two solutions, namely, x = 4 and x = 9. The gcd(4, 10) = 2 and 2 does not divide 5, but does divide 6.

Since gcd(3, 10) = 1, the linear congruence 3x ≡ 1 (mod 10) will have solutions, that is, modular multiplicative inverses of 3 modulo 10 will exist. In fact, 7 satisfies this congruence (i.e., 21 − 1 = 20). However, other integers also satisfy the congruence, for instance 17 and −3 (i.e., 3(17) − 1 = 50 and 3(−3) − 1 = −10). In particular, every integer in will satisfy the congruence since these integers have the form 7 + 10r for some integer r and

is divisible by 10. This congruence has only this one congruence class of solutions. The solution in this case could have been obtained by checking all possible cases, but systematic algorithms would be needed for larger moduli and these will be given in the next section.

The product of congruence classes and can be obtained by selecting an element of , say 25, and an element of , say −2, and observing that their product (25)(−2) = −50 is in the congruence class . Thus, . Addition is defined in a similar way. The ten congruence classes together with these operations of addition and multiplication of congruence classes form the ring of integers modulo 10, i.e., .

A complete residue system modulo 10 can be the set {10, −9, 2, 13, 24, −15, 26, 37, 8, 9} where each integer is in a different congruence class modulo 10. The unique least residue system modulo 10 is {0, 1, 2, ..., 9}. A reduced residue system modulo 10 could be {1, 3, 7, 9}. The product of any two congruence classes represented by these numbers is again one of these four congruence classes. This implies that these four congruence classes form a group, in this case the cyclic group of order four, having either 3 or 7 as a (multiplicative) generator. The represented congruence classes form the group of units of the ring . These congruence classes are precisely the ones which have modular multiplicative inverses.

Computation

Extended Euclidean algorithm

A modular multiplicative inverse of a modulo m can be found by using the extended Euclidean algorithm.

The Euclidean algorithm determines the greatest common divisor (gcd) of two integers, say a and m. If a has a multiplicative inverse modulo m, this gcd must be 1. The last of several equations produced by the algorithm may be solved for this gcd. Then, using a method called "back substitution", an expression connecting the original parameters and this gcd can be obtained. In other words, integers x and y can be found to satisfy Bézout's identity,

Rewritten, this is

that is,

so, a modular multiplicative inverse of a has been calculated. A more efficient version of the algorithm is the extended Euclidean algorithm, which, by using auxiliary equations, reduces two passes through the algorithm (back substitution can be thought of as passing through the algorithm in reverse) to just one.

In big O notation, this algorithm runs in time O(log2(m)), assuming |a| < m, and is considered to be very fast and generally more efficient than its alternative, exponentiation.

Using Euler's theorem

As an alternative to the extended Euclidean algorithm, Euler's theorem may be used to compute modular inverses. [11]

According to Euler's theorem, if a is coprime to m, that is, gcd(a, m) = 1, then

where is Euler's totient function. This follows from the fact that a belongs to the multiplicative group × if and only if a is coprime to m. Therefore, a modular multiplicative inverse can be found directly:

In the special case where m is a prime, and a modular inverse is given by

This method is generally slower than the extended Euclidean algorithm, but is sometimes used when an implementation for modular exponentiation is already available. Some disadvantages of this method include:

One notable advantage of this technique is that there are no conditional branches which depend on the value of a, and thus the value of a, which may be an important secret in public-key cryptography, can be protected from side-channel attacks. For this reason, the standard implementation of Curve25519 uses this technique to compute an inverse.

Multiple inverses

It is possible to compute the inverse of multiple numbers ai, modulo a common m, with a single invocation of the Euclidean algorithm and three multiplications per additional input. [12] The basic idea is to form the product of all the ai, invert that, then multiply by aj for all ji to leave only the desired a−1
i
.

More specifically, the algorithm is (all arithmetic performed modulo m):

  1. Compute the prefix products for all in.
  2. Compute b−1
    n
    using any available algorithm.
  3. For i from n down to 2, compute
    • a−1
      i
      = b−1
      i
      bi−1
      and
    • b−1
      i−1
      = b−1
      i
      ai
      .
  4. Finally, a−1
    1
    = b−1
    1
    .

It is possible to perform the multiplications in a tree structure rather than linearly to exploit parallel computing.

Applications

Finding a modular multiplicative inverse has many applications in algorithms that rely on the theory of modular arithmetic. For instance, in cryptography the use of modular arithmetic permits some operations to be carried out more quickly and with fewer storage requirements, while other operations become more difficult. [13] Both of these features can be used to advantage. In particular, in the RSA algorithm, encrypting and decrypting a message is done using a pair of numbers that are multiplicative inverses with respect to a carefully selected modulus. One of these numbers is made public and can be used in a rapid encryption procedure, while the other, used in the decryption procedure, is kept hidden. Determining the hidden number from the public number is considered to be computationally infeasible and this is what makes the system work to ensure privacy. [14]

As another example in a different context, consider the exact division problem in computer science where you have a list of odd word-sized numbers each divisible by k and you wish to divide them all by k. One solution is as follows:

  1. Use the extended Euclidean algorithm to compute k−1, the modular multiplicative inverse of k mod 2w, where w is the number of bits in a word. This inverse will exist since the numbers are odd and the modulus has no odd factors.
  2. For each number in the list, multiply it by k−1 and take the least significant word of the result.

On many machines, particularly those without hardware support for division, division is a slower operation than multiplication, so this approach can yield a considerable speedup. The first step is relatively slow but only needs to be done once.

Modular multiplicative inverses are used to obtain a solution of a system of linear congruences that is guaranteed by the Chinese Remainder Theorem.

For example, the system

X ≡ 4 (mod 5)
X ≡ 4 (mod 7)
X ≡ 6 (mod 11)

has common solutions since 5,7 and 11 are pairwise coprime. A solution is given by

X = t1 (7 × 11) × 4 + t2 (5 × 11) × 4 + t3 (5 × 7) × 6

where

t1 = 3 is the modular multiplicative inverse of 7 × 11 (mod 5),
t2 = 6 is the modular multiplicative inverse of 5 × 11 (mod 7) and
t3 = 6 is the modular multiplicative inverse of 5 × 7 (mod 11).

Thus,

X = 3 × (7 × 11) × 4 + 6 × (5 × 11) × 4 + 6 × (5 × 7) × 6 = 3504

and in its unique reduced form

X ≡ 3504 ≡ 39 (mod 385)

since 385 is the LCM of 5,7 and 11.

Also, the modular multiplicative inverse figures prominently in the definition of the Kloosterman sum.

See also

Notes

  1. Rosen 1993 , p. 132.
  2. Schumacher 1996 , p. 88.
  3. Stinson, Douglas R. (1995), Cryptography / Theory and Practice, CRC Press, pp. 124–128, ISBN   0-8493-8521-0
  4. Trappe & Washington 2006 , pp. 164−169.
  5. Moriarty, K.; Kaliski, B.; Jonsson, J.; Rusch, A. (2016). "PKCS #1: RSA Cryptography Specifications Version 2.2". Internet Engineering Task Force RFC 8017 . Internet Engineering Task Force. Retrieved January 21, 2017.
  6. Other notations are often used, including [a] and [a]m.
  7. Ireland & Rosen 1990 , p. 32
  8. Shoup, Victor (2005), A Computational Introduction to Number Theory and Algebra, Cambridge University Press, Theorem 2.4, p. 15, ISBN   9780521851541
  9. Rosen 1993 , p. 121
  10. Ireland & Rosen 1990 , p. 31
  11. Thomas Koshy. Elementary number theory with applications, 2nd edition. ISBN   978-0-12-372487-8. P. 346.
  12. Brent, Richard P.; Zimmermann, Paul (December 2010). "§2.5.1 Several inversions at once" (PDF). Modern Computer Arithmetic. Cambridge Monographs on Computational and Applied Mathematics. Vol. 18. Cambridge University Press. pp. 67–68. ISBN   978-0-521-19469-3.
  13. Trappe & Washington 2006 , p. 167
  14. Trappe & Washington 2006 , p. 165

Related Research Articles

<span class="mw-page-title-main">Chinese remainder theorem</span> Theorem for solving simultaneous congruences

In mathematics, the Chinese remainder theorem states that if one knows the remainders of the Euclidean division of an integer n by several integers, then one can determine uniquely the remainder of the division of n by the product of these integers, under the condition that the divisors are pairwise coprime.

<span class="mw-page-title-main">Modular arithmetic</span> Computation modulo a fixed integer

In mathematics, modular arithmetic is a system of arithmetic for integers, where numbers "wrap around" when reaching a certain value, called the modulus. The modern approach to modular arithmetic was developed by Carl Friedrich Gauss in his book Disquisitiones Arithmeticae, published in 1801.

<span class="mw-page-title-main">Quadratic reciprocity</span> Gives conditions for the solvability of quadratic equations modulo prime numbers

In number theory, the law of quadratic reciprocity is a theorem about modular arithmetic that gives conditions for the solvability of quadratic equations modulo prime numbers. Due to its subtlety, it has many formulations, but the most standard statement is:

<span class="mw-page-title-main">Gaussian integer</span> Complex number whose real and imaginary parts are both integers

In number theory, a Gaussian integer is a complex number whose real and imaginary parts are both integers. The Gaussian integers, with ordinary addition and multiplication of complex numbers, form an integral domain, usually written as or

The Jacobi symbol is a generalization of the Legendre symbol. Introduced by Jacobi in 1837, it is of theoretical interest in modular arithmetic and other branches of number theory, but its main use is in computational number theory, especially primality testing and integer factorization; these in turn are important in cryptography.

In number theory, Euler's criterion is a formula for determining whether an integer is a quadratic residue modulo a prime. Precisely,

In modular arithmetic, a number g is a primitive root modulo n if every number a coprime to n is congruent to a power of g modulo n. That is, g is a primitive root modulo n if for every integer a coprime to n, there is some integer k for which gka. Such a value k is called the index or discrete logarithm of a to the base g modulo n. So g is a primitive root modulo n if and only if g is a generator of the multiplicative group of integers modulo n.

In number theory, an integer q is called a quadratic residue modulo n if it is congruent to a perfect square modulo n; i.e., if there exists an integer x such that:

In algebra and number theory, Wilson's theorem states that a natural number n > 1 is a prime number if and only if the product of all the positive integers less than n is one less than a multiple of n. That is, the factorial satisfies

The quadratic sieve algorithm (QS) is an integer factorization algorithm and, in practice, the second fastest method known. It is still the fastest for integers under 100 decimal digits or so, and is considerably simpler than the number field sieve. It is a general-purpose factorization algorithm, meaning that its running time depends solely on the size of the integer to be factored, and not on special structure or properties. It was invented by Carl Pomerance in 1981 as an improvement to Schroeppel's linear sieve.

Modular exponentiation is exponentiation performed over a modulus. It is useful in computer science, especially in the field of public-key cryptography, where it is used in both Diffie-Hellman Key Exchange and RSA public/private keys.

In modular arithmetic computation, Montgomery modular multiplication, more commonly referred to as Montgomery multiplication, is a method for performing fast modular multiplication. It was introduced in 1985 by the American mathematician Peter L. Montgomery.

In number theory, Dixon's factorization method is a general-purpose integer factorization algorithm; it is the prototypical factor base method. Unlike for other factor base methods, its run-time bound comes with a rigorous proof that does not rely on conjectures about the smoothness properties of the values taken by a polynomial.

Multiplicative group of integers modulo <i>n</i> Group of units of the ring of integers modulo n

In modular arithmetic, the integers coprime to n from the set of n non-negative integers form a group under multiplication modulo n, called the multiplicative group of integers modulo n. Equivalently, the elements of this group can be thought of as the congruence classes, also known as residues modulo n, that are coprime to n. Hence another name is the group of primitive residue classes modulo n. In the theory of rings, a branch of abstract algebra, it is described as the group of units of the ring of integers modulo n. Here units refers to elements with a multiplicative inverse, which, in this ring, are exactly those coprime to n.

In mathematics, Hensel's lemma, also known as Hensel's lifting lemma, named after Kurt Hensel, is a result in modular arithmetic, stating that if a univariate polynomial has a simple root modulo a prime number p, then this root can be lifted to a unique root modulo any higher power of p. More generally, if a polynomial factors modulo p into two coprime polynomials, this factorization can be lifted to a factorization modulo any higher power of p.

The Tonelli–Shanks algorithm is used in modular arithmetic to solve for r in a congruence of the form r2n, where p is a prime: that is, to find a square root of n modulo p.

Cubic reciprocity is a collection of theorems in elementary and algebraic number theory that state conditions under which the congruence x3 ≡ p (mod q) is solvable; the word "reciprocity" comes from the form of the main theorem, which states that if p and q are primary numbers in the ring of Eisenstein integers, both coprime to 3, the congruence x3p is solvable if and only if x3q is solvable.

In mathematics, elliptic curve primality testing techniques, or elliptic curve primality proving (ECPP), are among the quickest and most widely used methods in primality proving. It is an idea put forward by Shafi Goldwasser and Joe Kilian in 1986 and turned into an algorithm by A. O. L. Atkin the same year. The algorithm was altered and improved by several collaborators subsequently, and notably by Atkin and François Morain, in 1993. The concept of using elliptic curves in factorization had been developed by H. W. Lenstra in 1985, and the implications for its use in primality testing followed quickly.

In number theory, a kth root of unity modulo n for positive integers k, n ≥ 2, is a root of unity in the ring of integers modulo n; that is, a solution x to the equation . If k is the smallest such exponent for x, then x is called a primitive kth root of unity modulo n. See modular arithmetic for notation and terminology.

In number theory, Berlekamp's root finding algorithm, also called the Berlekamp–Rabin algorithm, is the probabilistic method of finding roots of polynomials over a field . The method was discovered by Elwyn Berlekamp in 1970 as an auxiliary to the algorithm for polynomial factorization over finite fields. The algorithm was later modified by Rabin for arbitrary finite fields in 1979. The method was also independently discovered before Berlekamp by other researchers.

References