Fact-checked by Grok 2 weeks ago

Modular multiplicative inverse

In modular arithmetic, the modular multiplicative inverse of an integer a modulo an integer m > 1 is an integer b such that a b \equiv 1 \pmod{m}, meaning m divides a b - 1 but not necessarily a or b. Such an inverse exists if and only if a and m are relatively prime, that is, their greatest common divisor satisfies \gcd(a, m) = 1. When it exists, the inverse is unique modulo m, and it enables "division" in the ring of integers modulo m by allowing multiplication by b to undo multiplication by a. The existence of modular multiplicative inverses is a cornerstone of elementary , directly tied to , which states that \gcd(a, m) = 1 there exist integers x and y such that a x + m y = 1; here, x \pmod{m} is the of a. Computationally, the is efficiently found using the , which not only computes \gcd(a, m) but also yields the coefficients from . For prime moduli p, every nonzero a modulo p has an , forming the \mathbb{Z}_p^\times, and provides a method to compute it as a^{p-2} \pmod{p}. Modular inverses play a vital role in applications beyond , particularly in , where they are essential for constructing secure systems like the algorithm. In , the private decryption exponent d is the modular multiplicative inverse of the public encryption exponent e modulo \phi(n), where n is the product of two large primes and \phi is , allowing efficient decryption while keeping the factorization of n secret. They also underpin and other protocols by facilitating and point decompression in finite fields.

Modular Arithmetic Basics

Congruences and Equivalence Classes

In modular arithmetic, two integers a and b are said to be congruent modulo n, where n is a positive integer, denoted a \equiv b \pmod{n}, if n divides the difference a - b. This relation is an on the set of integers, satisfying reflexivity (a \equiv a \pmod{n}), (if a \equiv b \pmod{n}, then b \equiv a \pmod{n}), and (if a \equiv b \pmod{n} and b \equiv c \pmod{n}, then a \equiv c \pmod{n}). The partitions the integers into equivalence classes, known as residue classes n. The residue class of a n, denoted _n, consists of all integers b such that b \equiv a \pmod{n}, or equivalently, _n = \{a + kn \mid k \in \mathbb{Z}\}. There are exactly n distinct residue classes n, typically represented by the integers $0, 1, \dots, n-1. Congruences support basic arithmetic operations: if a \equiv b \pmod{n} and c \equiv d \pmod{n}, then a + c \equiv b + d \pmod{n} and a \cdot c \equiv b \cdot d \pmod{n}. These operations are well-defined on the set of residue classes, allowing arithmetic to be performed n without altering . The set of residue classes n, denoted \mathbb{Z}/n\mathbb{Z}, forms a under these addition and multiplication operations, with the $0 \pmod{n} and multiplicative identity $1 \pmod{n}.

Greatest Common Divisor in Modular Contexts

In modular arithmetic, the greatest common divisor (GCD) of two integers a and n, where n > 0, plays a crucial role in determining the structure of solutions to equations within the ring of integers modulo n. The Euclidean algorithm provides an efficient method to compute \gcd(a, n) by leveraging the division algorithm repeatedly. Specifically, the algorithm states that \gcd(a, n) = \gcd(n, a \mod n), and it proceeds by replacing a with n and n with the remainder a \mod n until the remainder is zero; the last non-zero remainder is the GCD. This process terminates because the remainders form a strictly decreasing sequence of non-negative integers. If the algorithm terminates with a GCD of 1, then a and n are coprime. For an illustrative example, consider computing \gcd(15, 28): \begin{align*} 28 &= 1 \cdot 15 + 13, \\ 15 &= 1 \cdot 13 + 2, \\ 13 &= 6 \cdot 2 + 1, \\ 2 &= 2 \cdot 1 + 0. \end{align*} The last non-zero is 1, so \gcd(15, 28) = 1, indicating that 15 and 28 are coprime. further connects the GCD to linear combinations: if d = \gcd(a, n), then there exist integers x and y such that a x + n y = d. Moreover, the set of all integer s of a and n consists precisely of the multiples of d. A proof sketch relies on the algorithm's steps, using back-substitution to express the GCD as such a combination. Starting from the base case where the is d (with previous 0), substitute backwards through the division steps. For instance, in the example above for \gcd(15, 28) = 1: \begin{align*} 1 &= 13 - 6 \cdot 2, \\ 2 &= 15 - 1 \cdot 13 \quad \Rightarrow \quad 1 = 13 - 6 \cdot (15 - 1 \cdot 13) = 7 \cdot 13 - 6 \cdot 15, \\ 13 &= 28 - 1 \cdot 15 \quad \Rightarrow \quad 1 = 7 \cdot (28 - 1 \cdot 15) - 6 \cdot 15 = 7 \cdot 28 - 13 \cdot 15. \end{align*} Thus, x = -13 and y = 7 satisfy $15(-13) + 28(7) = 1. This back-substitution works generally because each is a linear combination of the originals, propagating the expression upwards. Two integers a and n are defined as coprime if \gcd(a, n) = 1. In the context of , which equate integers differing by multiples of the , coprimality has significant implications for solvability. Specifically, the linear a x \equiv b \pmod{n} is solvable for x \gcd(a, n) divides b; when \gcd(a, n) = 1, the is always solvable for any b, with exactly one solution n. This condition ensures that the coefficient a generates all residue classes n when coprime, facilitating unique solutions in modular equations.

Definition and Conditions

Formal Definition

In modular arithmetic, the modular multiplicative inverse of an integer a modulo n, where n > 1 is a positive integer, is an integer x satisfying the congruence a x \equiv 1 \pmod{n}. This x is commonly denoted as x \equiv a^{-1} \pmod{n}. Algebraically, this concept corresponds to the in the \mathbb{Z}/n\mathbb{Z} of integers n, where a and its x are elements whose product is the multiplicative $1 in the . The is symmetric: if x is the modular multiplicative of a n, then a is the modular multiplicative of x n.

Existence Criteria

The existence of a modular multiplicative inverse for an a n > 1 is determined by a fundamental condition in : such an inverse exists \gcd(a, n) = 1. This criterion ensures that a and n are coprime, meaning they share no common prime factors other than 1. When this holds, there exists an x such that a x \equiv 1 \pmod{n}. To establish this theorem, consider the necessity first. Suppose an inverse x exists, so a x \equiv 1 \pmod{n}, which implies a x - n k = 1 for some integer k. Any common divisor of a and n must divide the left side a x - n k, hence it divides 1, forcing \gcd(a, n) = 1. For sufficiency, if \gcd(a, n) = 1, guarantees integers x and y such that a x + n y = 1, or equivalently, a x \equiv 1 \pmod{n}. Thus, x serves as the inverse. If \gcd(a, n) = d > 1, no inverse exists. In this case, d divides a, so for any x, a x \equiv 0 \pmod{d}. However, since d does not divide 1, a x \not\equiv 1 \pmod{d}, and thus a x \not\equiv 1 \pmod{n} because n is a multiple of d. This impossibility arises because the equation a x \equiv 1 \pmod{n} would require consistency modulo every of n, including d. This existence criterion extends to the broader solvability of linear congruences. The equation a x \equiv b \pmod{n} has solutions \gcd(a, n) divides b. When b = 1, this reduces to the case, but the allows for solutions even when \gcd(a, n) > 1, provided the gcd divides b, with the number of distinct solutions n then equaling the gcd value.

Properties and Uniqueness

Uniqueness of Inverses

When a modular multiplicative inverse exists for an a modulo n, it is unique modulo n. Specifically, if ax \equiv 1 \pmod{n} and ay \equiv 1 \pmod{n}, then x \equiv y \pmod{n}. To prove this, subtract the congruences to obtain a(x - y) \equiv 0 \pmod{n}, meaning n divides a(x - y). Since the existence of the inverse requires \gcd(a, n) = 1, it follows that n divides (x - y), so x \equiv y \pmod{n}. This uniqueness holds under the existence criterion that \gcd(a, n) = 1. The modular inverses modulo n exhibit several basic properties. First, the set of integers coprime to n (those possessing inverses) is closed under modulo n: if \gcd(a, n) = 1 and \gcd(b, n) = 1, then \gcd(ab \mod n, n) = 1. Second, the inverse of an inverse recovers the original element: if b is the inverse of a modulo n, then the inverse of b is a modulo n. This follows because ab \equiv 1 \pmod{n} and, by commutativity of , ba \equiv 1 \pmod{n}, so a satisfies the defining for the inverse of b. Finally, the of the -a n is the of the of a: if b is the of a, then -b \mod n is the of -a \mod n, since (-a)(-b) \equiv ab \equiv 1 \pmod{n}.

Connection to Units in Rings

In the ring \mathbb{Z}/n\mathbb{Z}, the units are the residue classes $$ where \gcd(a, n) = 1, as these are precisely the elements that admit multiplicative inverses under the ring's operation. These units form a multiplicative known as the group of units, denoted U(n) or (\mathbb{Z}/n\mathbb{Z})^\times, consisting of all such invertible elements closed under modulo n. The group U(n) is abelian, reflecting the commutative nature of multiplication in \mathbb{Z}/n\mathbb{Z}. Its order is given by \phi(n), which counts the number of integers from 1 to n-1 coprime to n. Within this group structure, the modular multiplicative inverse of an element \in U(n) coincides with its group inverse [a^{-1}], satisfying \cdot [a^{-1}] = {{grok:render&&&type=render_inline_citation&&&citation_id=1&&&citation_type=wikipedia}} in \mathbb{Z}/n\mathbb{Z}. As a consequence of , this inverse is unique for each unit. A key structural property arises from the : when n = pq with \gcd(p, q) = 1, the ring \mathbb{Z}/n\mathbb{Z} \cong \mathbb{Z}/p\mathbb{Z} \times \mathbb{Z}/q\mathbb{Z} induces a U(n) \cong U(p) \times U(q). This decomposition allows inverses modulo n to be computed via separate inverses modulo p and q, combined using the theorem's reconstruction mechanism.

Computation Techniques

Extended Euclidean Algorithm

The extended Euclidean algorithm provides an efficient method to compute the modular multiplicative inverse of an integer a modulo n, assuming \gcd(a, n) = 1. This algorithm extends the standard for finding the by also determining integers x and y such that a x + n y = 1, based on . When \gcd(a, n) = 1, the value x \mod n serves as the modular inverse, since a x \equiv 1 \pmod{n}. The algorithm proceeds in two phases. First, apply the to compute \gcd(a, n) through successive divisions: divide n by a to get q_1 and r_1 = n - q_1 a, then a by r_1 to get q_2 and r_2 = a - q_2 r_1, and continue until the remainder is 1 (since the gcd is 1). This generates a sequence of equations expressing each as a of a and n. In the second phase, back-substitute starting from $1 = r_k (the last non-zero ) upwards, substituting previous expressions to solve for x and y. The coefficients are updated iteratively: initialize s_0 = 1, s_1 = 0, then for each step i \geq 2, s_i = s_{i-2} - q_i s_{i-1}, where q_i is the at that step; the final s_k gives x. Consider the example of finding the inverse of 5 17. Apply the :
$17 = 3 \cdot 5 + 2,
$5 = 2 \cdot 2 + 1,
$2 = 2 \cdot 1 + 0.
Thus, \gcd(5, 17) = 1. Now back-substitute:
$1 = 5 - 2 \cdot 2,
but $2 = 17 - 3 \cdot 5, so
$1 = 5 - 2 \cdot (17 - 3 \cdot 5) = 5 - 2 \cdot 17 + 6 \cdot 5 = 7 \cdot 5 - 2 \cdot 17.
Hence, x = 7 and y = -2, and $5 \cdot 7 = 35 \equiv 1 \pmod{17}, confirming 7 as the inverse.
The following pseudocode implements the recursive version, returning the gcd d, x, and y:
function extended_gcd(a, n):
    if n == 0:
        return a, 1, 0
    d, s1, t1 = extended_gcd(n, a % n)
    s = t1
    t = s1 - (a // n) * t1
    return d, s, t
To compute the inverse, call extended_gcd(a, n); if d = 1, the inverse is s \mod n. The time complexity of the extended Euclidean algorithm is O(\log \min(a, n)), matching the standard Euclidean algorithm due to the logarithmic number of division steps.

Euler's Theorem Method

Euler's theorem provides a method for computing the modular multiplicative inverse of an integer a modulo n when \gcd(a, n) = 1. The theorem states that if a and n are coprime, then a^{\phi(n)} \equiv 1 \pmod{n}, where \phi(n) is Euler's totient function. From this congruence, it follows that a^{\phi(n) - 1} \equiv a^{-1} \pmod{n}, since multiplying both sides by a^{-1} (which exists under the coprimality condition) yields the identity. To apply this method, first compute \phi(n), defined as \phi(n) = n \prod_{p \mid n} \left(1 - \frac{1}{p}\right), where the product is over the distinct prime factors p of n. Then, calculate a^{\phi(n) - 1} \mod n using efficient algorithms, such as binary exponentiation, which reduces the computational complexity to O(\log(\phi(n))) multiplications modulo n. This approach leverages the group structure of the units modulo n, whose is \phi(n), but focuses practically on exponentiation rather than group-theoretic operations. For example, consider finding the of 3 modulo 10. Since \gcd(3, 10) = 1 and \phi(10) = 10 \left(1 - \frac{1}{2}\right)\left(1 - \frac{1}{5}\right) = 4, compute $3^{4-1} = 3^3 = 27 \equiv 7 \pmod{10}. Verifying, $3 \times 7 = 21 \equiv 1 \pmod{10}, confirming that 7 is the . While theoretically elegant due to its connection to the totient function, this method can be inefficient for large n without optimized , as \phi(n) may still be substantial and require of n upfront. It remains valuable for theoretical analysis and cases where is preferred over other techniques.

Illustrative Examples

Basic Numerical Examples

To illustrate the concept of a modular multiplicative inverse, consider the case of finding the inverse of 7 26. Since \gcd(7, 26) = 1, an inverse exists, and using the yields 15 as the solution, as $7 \times 15 = 105 \equiv 1 \pmod{26}. To verify, compute the product and reduce: $105 \div 26 = 4 1, confirming $105 \equiv 1 \pmod{26}. In contrast, the 4 has no modular multiplicative inverse 6, because \gcd(4, 6) = 2 \neq 1. This violates the existence condition, rendering the $4x \equiv 1 \pmod{6} unsolvable for any x, as the does not divide 1. For small , the inverses of numbers coprime to the can be tabulated systematically. The following table lists the modular multiplicative inverses 5 (where \gcd(a, 5) = 1 for a = 1, 2, 3, 4):
aa^{-1} \pmod{5}Verification
11$1 \times 1 = 1 \equiv 1 \pmod{5}
2$2 \times 3 = 6 \equiv 1 \pmod{5}
2$3 \times 2 = 6 \equiv 1 \pmod{5}
44$4 \times 4 = 16 \equiv 1 \pmod{5}
These pairs are derived from the multiplication table modulo 5, highlighting the symmetry where distinct inverses are mutual.

Advanced Applications in Coding

In Reed-Solomon codes, defined over finite GF(2^m), modular multiplicative inverses play a crucial role in decoding processes, particularly for computing syndromes and correcting errors. These codes treat as evaluated at elements of the , with the having roots at consecutive powers of a primitive element α. During decoding, syndromes are calculated as S_i = R(α^i), where R(x) is the received ; subsequent steps, such as determining error locations and values, require divisions, which are implemented via multiplication by inverses. In GF(2^m), inverses are computed efficiently using the applied to modulo the 's irreducible , ensuring operations remain within the structure. A historical advancement in this area is the Berlekamp-Massey algorithm, developed in 1969, which iteratively finds the shortest matching a sequence of syndromes, relying on multiplicative inverses to update coefficients and normalize the error-locator polynomial σ(x). This algorithm is essential for efficient decoding of Reed-Solomon codes, as it locates multiple errors in O(t^2) time, where t is the error-correcting capability, with inverses ensuring precise coefficient adjustments during iterations. Consider a practical example in GF(16), constructed with the x^4 + x + 1 = 0 over GF(2), where α is a primitive element satisfying α^4 = α + 1. For a (15,13) Reed-Solomon code capable of correcting a single (t=1), suppose the received R(x) yields syndromes S_1 = α^3 and S_2 = α^4 after evaluation at α and α^2. The location X_1 is given by X_1 = S_2 \cdot S_1^{-1}, and the value Y_1 = S_1 \cdot X_1^{-1}. To compute S_1^{-1} = (α^3)^{-1}, apply the to the representation of α^3 (which is x^3 the irreducible) and the field modulus x^4 + x + 1. The steps involve repeated and back-substitution: first, divide x^4 + x + 1 by x^3 to get x and x + 1; continue until the GCD is 1, yielding coefficients s and t such that s \cdot (α^3) + t \cdot (α^4 + α + 1) = 1, so s(α) = (α^3)^{-1} = α^{12}. With X_1 = α^4 \cdot α^{12} = α^1 and Y_1 computed similarly as α^2, the decoder subtracts Y_1 from R(x) at position X_1, correcting the single and recovering the original codeword. This demonstrates how inverses enable targeted correction in extensions.

Practical Applications

Cryptography and Security

The modular multiplicative inverse is fundamental to the RSA cryptosystem, introduced in 1978, where it enables the generation of the private key from the public key components. In RSA, a user selects two large primes p and q, computes the modulus n = p \times q, and chooses a public exponent e coprime to \phi(n) = (p-1)(q-1). The private exponent d is then the modular inverse of e modulo \phi(n), satisfying e \times d \equiv 1 \pmod{\phi(n)}. This inverse allows decryption of a ciphertext c (obtained as m^e \mod n for plaintext m) via m = c^d \mod n. For illustration, consider primes p=3 and q=11, so n=33 and \phi(33)=20. With public exponent e=3 (coprime to 20), the private exponent is d=7 since $3 \times 7 = 21 \equiv 1 \pmod{20}. To encrypt plaintext m=5, compute ciphertext c = 5^3 \mod 33 = 125 \mod 33 = 26. Decryption yields $26^7 \mod 33 = 5, recovering the original message. This process demonstrates how the inverse ensures reversible encryption in the ring of integers modulo n. Modular multiplicative inverses also underpin the Diffie-Hellman key exchange protocol from 1976, operating in the multiplicative group modulo a large prime p. Here, parties agree on p and a generator g, then exchange g^a \mod p and g^b \mod p (for private exponents a and b); each computes the shared secret (g^b)^a \mod p = g^{ab} \mod p or symmetrically. Inverses exist for all nonzero elements modulo p (a field), enabling group operations essential for secure key derivation without direct transmission. The protocol's security stems from the discrete logarithm problem's hardness in this setting. RSA's security relies on the presumed difficulty of factoring n to recover \phi(n) and compute the private inverse d, as inverting e modulo \phi(n) without factorization is computationally infeasible for large n. Without the trapdoor (knowledge of p and q), deriving d requires solving the difficult problem of finding Euler's totient or directly inverting in an unknown modulus. However, Shor's , proposed in 1994, factors n in time using a quantum computer, potentially breaking RSA by enabling inverse computation. Post-2023 advancements, including optimized implementations, have lowered the barrier: a 2025 Quantum AI analysis estimates RSA-2048 could be factored in under a week with fewer than one million noisy qubits, accelerating the shift to .

Error-Correcting Codes

In linear error-correcting codes defined over the \mathbb{Z}/p\mathbb{Z} (where p is prime), modular multiplicative inverses play a fundamental role by enabling the solution of linear systems required for . Specifically, these codes operate as subspaces of (\mathbb{Z}/p\mathbb{Z})^n, and decoding often involves computing the s = H e (where H is the parity-check and e is the ) and solving for e through operations that rely on inverses to isolate error locations and values. This structure ensures that every nonzero element has a unique , facilitating efficient algebraic manipulations essential for practical implementations. Hamming codes exemplify this application, as they are perfect linear codes capable of correcting single over finite fields \mathbb{F}_p. In the binary Hamming(7,4) over \mathbb{GF}(2), the parity-check H consists of all nonzero columns from \mathbb{F}_2^3, and the directly identifies the position since field inverses are trivial (the inverse of 1 is 1 2). For extensions to odd primes p > 2, such as Hamming codes over \mathbb{F}_p, decoding requires explicit computation of inverses to resolve parity-check equations; for instance, solving H e = s may involve premultiplying by H^{-1} or , where inverses of pivot elements p determine the e. This generalization enhances performance in non-binary settings, allowing correction of in symbols from larger alphabets. Advancements in low-density parity-check (LDPC) codes further highlight the utility of modular inverses in iterative decoding over finite fields \mathbb{GF}(q). In algorithms like the sum-product algorithm (), check-node and variable-node updates involve convolutions and normalizations over \mathbb{GF}(q), where multiplicative inverses are invoked to compute probabilities or messages (e.g., factors in log-domain implementations using fast transforms over extension fields). Simplified variants, such as the FFT-based and min-max approximations, reduce complexity by a factor of q while preserving performance close to the full , relying on efficient inverse computations for field multiplications during . These methods achieve near-capacity error correction in high-rate scenarios. Reed-Solomon codes, widely used in storage media (e.g., CDs, DVDs) and digital communications (e.g., QR codes, satellite links), also rely on modular multiplicative inverses in finite fields \mathbb{GF}(2^m). These non-binary cyclic codes correct multiple symbol errors through decoding algorithms like Berlekamp-Massey, which solve key equations using field inverses to find error locator polynomials. In research for beyond-5G systems like 6G, non-binary variants of polar and LDPC codes are proposed for enhanced reliability in control channels, incorporating inverses in polar transform operations and rate-matching to optimize error correction under varying channel conditions.

References

  1. [1]
    [PDF] Everything You Need to Know About Modular Arithmetic...
    Definition An inverse to a modulo m is a integer b such that ... Then a has a multiplicative inverse modulo m if a and m are relatively prime.Missing: theory | Show results with:theory
  2. [2]
    Math 511, Modular Inverses
    We write the greatest common divisor of a and n as gcd(a,n). With this notation, we conjecture that a has an inverse (mod n) if and only if gcd(a,n)=1.Missing: theory | Show results with:theory
  3. [3]
    [PDF] Number Theory
    Theorem 5.17 (Existance of modular multiplicative inverse) Consider some integers a and m > 1. If gcd(a, m) = 1, then a has a unique modular multiplicative ...
  4. [4]
    3.4 Multiplicative Inverses
    Any number that has a 1 in its row/column has an inverse. These are 1, 2, 4, 5, 7, 8. The other numbers: 3, 6 (and 0) do not have multiplicative inverses.
  5. [5]
    Number Theory - Modular Arithmetic
    We consider two integers x , y to be the same if x and y differ by a multiple of n , and we write this as x = y ( mod n ) , and say that x and y are ...
  6. [6]
    [PDF] a3 integers 23 (2023) pairwise modular multiplicative inverses and ...
    Jan 13, 2023 · Abstract. Let (p, q) be a pair of relatively prime integers greater than 1. The pairwise modular multiplicative inverse, or PMMI for short, ...
  7. [7]
    [PDF] Number Theory and RSA Public-Key Encryption
    The decryption key d is the multiplicative inverse of 11 modulo 120. • Run the Extended Euclid algorithm with m = 120 and n = 11. • We find the decryption key d ...
  8. [8]
    Modular Inverse for Integers using Fast Constant Time GCD ...
    Modular inversion, the multiplicative inverse of an integer in the ring of integers modulo a prime number, is widely used in public-key cryptography.
  9. [9]
    [PDF] Basic Properties of Congruences
    The letters a, b, c, d, k represent integers. The letters m, n represent positive integers. The notation a ≡ b (mod m) means that m divides a − b.
  10. [10]
    [PDF] MTHSC 412 Section 2.5 –Congruence of Integers
    Thus x ≡ z (mod n) and ≡ (mod n) is transitive. Since ≡ (mod n) is reflexive, symmetric and transitive, it is an equivalence relation on Z . Kevin ...<|control11|><|separator|>
  11. [11]
    NTIC Properties of Congruence
    Addition and multiplication modulo n n are well-defined . That is, if a≡c a ≡ c and b≡d b ≡ d (modulo some fixed modulus n n ), then both of these ...
  12. [12]
    [PDF] Ring Theory - The University of Memphis
    2. Z/nZ is a ring under + and × mod n. This ring is an ID iff n is prime. In fact, if n is prime then Z/nZ is a field.<|control11|><|separator|>
  13. [13]
    The Euclidean Algorithm (article) - Khan Academy
    The Euclidean Algorithm is a technique for quickly finding the GCD of two integers. The Algorithm The Euclidean Algorithm for finding GCD(A,B) is as follows:
  14. [14]
    Euclidean algorithm for computing the greatest common divisor
    Oct 15, 2024 · The Euclidean algorithm was formulated as follows: subtract the smaller number from the larger one until one of the numbers is zero.Algorithm · Implementation · Time Complexity · Least common multiple
  15. [15]
    Methods to Find GCF of 15 and 28 - Cuemath
    There are 3 commonly used methods to find the GCF of 15 and 28 - prime factorization, Euclidean algorithm, and long division.
  16. [16]
  17. [17]
    Bezout's Identity | Brilliant Math & Science Wiki
    By the Euclidean algorithm, there must be integers ℓ \ell ℓ and m m m such ... The proof of Bézout's identity uses the property that for nonzero ...
  18. [18]
    5.4 Linear congruences | MATH1001 Introduction to Number Theory
    If gcd(a,n)=1 gcd ( a , n ) = 1 then the solutions of the linear congruence ax≡b mod (n) a x ≡ b mod ( n ) form a single congruence class mod (n) mod ( n ) .Missing: coprimality solvability
  19. [19]
    Modular Inverse -- from Wolfram MathWorld
    A modular inverse of an integer b (modulo m ) is the integer b^(-1) such that bb^(-1)=1 (mod m). A modular inverse can be computed in the Wolfram Language.
  20. [20]
    [PDF] Introduction to Modular Arithmetic∗ 1 Integers modulo n
    Let n be a positive integer. The equivalence classes under the relation of congruence modulo n are called the residue classes modulo n. A residue class modulo n.
  21. [21]
    [PDF] Lecture 10 - People @EECS
    Reducing b modulo m gives us the unique inverse we are looking for. In the above example, we see that 3 is the multiplicative inverse of 12 mod 35.
  22. [22]
    [PDF] Number Theory Course notes for MA 341, Spring 2018
    May 2, 2018 · You have already seen that an inverse is unique if it exists. Theorem 4.1.1. a has a multiplicative inverse modulo m if and only if gcd(a, m)=1.
  23. [23]
    [PDF] Math 43900 Problem Solving Fall 2021 Lecture 7 Number Theory
    Bézout's identity: If m and n are two integers with gcd d there exist integers a and b such that am+bn = d. In other words, m has a multiplicative inverse mod n ...
  24. [24]
    [PDF] 8.1 Primality testing using modular arithmetic - MIT Mathematics
    Jan 27, 2017 · since if gcd(a, n) > 1 then no multiple of a can be congruent to 1 modulo n (so [a] has no multiplicative inverse) and if gcd(a, n) = 1 then we ...<|control11|><|separator|>
  25. [25]
    [PDF] Contents 2 Modular Arithmetic in Z - Evan Dummit
    ◦ Proof: If x is a solution to the congruence ax ≡ b (mod m), then there exists an integer k with ax−mk = b. Since d = gcd(a, m) divides the left-hand side, it ...Missing: criteria | Show results with:criteria
  26. [26]
    [PDF] Modular arithmetic - Department of Mathematics | University of Toronto
    Let 𝑎 ∈ {1,2,…,𝑛 − 1} then gcd(𝑎,𝑛) = 1. Hence 𝑎 admits a multiplicative inverse modulo 𝑛, so there exists 𝑏 ∈ {1,2,…,𝑛 − 1} such that 𝑎𝑏 ≡ 1 (mod 𝑛).
  27. [27]
    None
    ### Summary of Uniqueness of Modular Inverse
  28. [28]
    [PDF] Computer Security Sources Modular Arithmetic Modular Arithmetic
    What does this mean for Z10*?. • It turns out that this is true for all n: Zn* is closed under multiplication mod n. 22. Fermat's Little Theorem. • *Fermat's ...
  29. [29]
    [PDF] SEVENTH LECTURE 1. The Unit Group of Z/nZ Consider a nonunit ...
    Let G be any finite subgroup of the unit group of any field. Then. G is cyclic. In particular, the multiplicative group modulo any prime p is cyclic,. (Z/pZ)× ∼.
  30. [30]
    Number Theory - Units and the Totient Function
    We know y is a unit if and only if y and n are coprime. So the size of Z n ∗ is precisely the number of integers in that are coprime to n.
  31. [31]
    [PDF] The Structure of (Z/nZ)
    Apr 6, 2018 · One can easily show that 2 + 11Z generates (Z/11Z)×. Since this group has order 10, the only other generators are (2 + 11Z)3 =8+11/Z, (2 + 11/Z ...
  32. [32]
    [PDF] Math 561/661 Fall 2023 Homework 3 Drew Armstrong - Mathematics
    We say that u ∈ Z/nZ is a unit if there exist some x ∈ Z/nZ such that ux ≡ 1 mod n. We denote the multiplicative group of units by ((Z/nZ)×,·,1). (a) Prove ...
  33. [33]
    [PDF] The Chinese Remainder Theorem
    Feb 19, 2018 · ρ : Z/NZ → Z/n1Z × Z/n2Z···× Z/n1Z a + NZ 7→ (a + Z/n1Z,a + Z/n2Z ... Z are units. This is easily explained. In order to be ...
  34. [34]
    [PDF] The Euclidean Algorithm and Multiplicative Inverses
    The Euclidean Algorithm is a set of instructions for finding the greatest common divisor of any two positive integers. Its original importance was probably ...
  35. [35]
    [PDF] CSE 311 Lecture 14: Euclidean Algorithm and Modular Equations
    So, we can compute multiplicative inverses with the extended Euclidean algorithm. These inverses let us solve modular equations. 12. Page 26. Modular equations.
  36. [36]
    None
    ### Summary of Extended Euclidean Algorithm from https://websites.nku.edu/~christensen/Euclidean%20algorithm%20and%20multiplicative%20inverses.pdf
  37. [37]
    Extended Euclid Algorithm - UMBC CSEE
    Here's the pseudo-code for the Extended Euclid algorithm: ExtEuclid (a,b) { // returns a triple (d,s,t) such that d = gcd(a,b) and // d == a*s + b*t if (b == 0 ...
  38. [38]
    Modular Inverse - Algorithms for Competitive Programming
    Aug 20, 2023 · In this article, we present two methods for finding the modular inverse in case it exists, and one method for finding the modular inverse for all numbers in ...
  39. [39]
    [PDF] (2) Modular multiplicative inverse Explanation Computation
    A modular multiplicative inverse of a modulo m is an integer x such that, and it can be found using the extended Euclidean algorithm.
  40. [40]
    Euler's Totient Theorem -- from Wolfram MathWorld
    A generalization of Fermat's little theorem. Euler published a proof of the following more general theorem in 1736. Let phi(n) denote the totient function.
  41. [41]
    Totient Function -- from Wolfram MathWorld
    The totient function phi(n), also called Euler's totient function, is defined as the number of positive integers <=n that are relatively prime to.
  42. [42]
    [PDF] EULER'S THEOREM 1. Introduction Fermat's little ... - Keith Conrad
    Introduction. Fermat's little theorem is an important property of integers to a prime modulus. Theorem 1.1 (Fermat). For prime p and any a ∈ Z such that a ...
  43. [43]
    7.1 Euler's Theorem | MATH1001 Introduction to Number Theory
    7.1 Euler's Theorem. Recall that a multiplicative inverse for a class [a]∈Zn [ a ] ∈ Z n is a class [b]∈Zn [ b ] ∈ Z n such that [a][b]=[1] [ a ] [ b ] ...
  44. [44]
    [PDF] CSE 311: Foundations of Computing - Washington
    We already computed that 15 is the multiplicative inverse of 7 modulo 26: That is, 7 · 15 ≡ 1 (mod 26). By the multiplicative property of mod we have.
  45. [45]
    [PDF] 8 Modular Arithmetic - Clemson University
    Now, we can write down tables for modular arithmetic. For example, here are the tables for arithmetic modulo 4 and modulo 5. +4. 0 1 2 3. 0 ...
  46. [46]
    [PDF] Tutorial on Reed-Solomon Error Correction Coding
    This tutorial covers Reed-Solomon error correction coding, including Reed-Solomon encoding, block codes, and error correction systems.
  47. [47]
    [PDF] An Introduction to Galois Fields and Reed-Solomon Coding
    Oct 4, 2010 · If p is prime then each element x also has a multiplicative inverse y whose value is chosen so that xy = 1 mod p. There is no simple formula for ...
  48. [48]
    [PDF] A Method for Obtaining Digital Signatures and Public-Key ...
    ITS ALL GREEK TO ME (Julius Caesar, I, ii, 288, paraphrased) is encoded: 0920 1900 0112 1200 0718 0505 1100 2015 0013 0500 Since e = 10001 in binary, the first ...
  49. [49]
    RSA - Stanford Computer Science
    RSA is a public-key cryptosystem using public and private keys. It uses Euler's theorem and the difficulty of factoring large numbers.<|control11|><|separator|>
  50. [50]
    [PDF] New Directions in Cryptography - Stanford Electrical Engineering
    Diffie and M. E. Hellman, “Multiuser cryptographic techniques,” presented at National Computer Conference, New York, June 7-10,. 1976. [6] D. Knuth, The Art of ...
  51. [51]
    Primes, Modular Arithmetic, and Public Key Cryptography
    Diffie-Hellman Key Exchange. The premise of the Diffie-Hellman key exchange is that two people, Alice and Bob, want to come up with a shared secret number.<|separator|>
  52. [52]
    Google Researcher Lowers Quantum Bar to Crack RSA Encryption
    May 24, 2025 · A new study from Google Quantum AI estimates that breaking RSA-2048 encryption could be achieved in under a week using fewer than one million ...
  53. [53]
    [PDF] error-correcting codes and finite fields
    Sep 29, 2012 · Abstract. We investigate the properties of modern error-correcting codes from an algebraic perspective. First, using techniques of linear ...Missing: modular | Show results with:modular
  54. [54]
    [PDF] Modular numbers and Error Correcting Codes • Introduction
    To see why the modular numbers Fp form a field, we must find a way to show that any nonzero modular number has an inverse. One easy way when the prime p is not ...
  55. [55]
    Decoding Algorithms for Nonbinary LDPC Codes Over GF<formula formulatype="inline"><tex>$(q)$</tex></formula>
    Insufficient relevant content. The provided content (title and metadata) does not contain detailed information about the paper's content on how modular multiplicative inverses are used in decoding nonbinary LDPC codes over GF(q). No full text or specific algorithmic details are available in the given input.