Fact-checked by Grok 2 weeks ago

Finite field

A finite field, also known as a Galois field, is an that consists of a of elements equipped with and operations satisfying the field axioms, including commutativity, associativity, distributivity, the existence of additive and multiplicative identities, additive inverses, and multiplicative inverses for nonzero elements. The number of elements in a finite field, referred to as its order, is always a power of a , denoted as p^n where p is a prime and n is a positive . For every prime p and positive n, there exists a finite field of p^n, and any two such fields are unique up to , meaning they have essentially the same structure regardless of how they are presented. This uniqueness follows from the fact that all finite fields of the same are isomorphic extensions of the prime field \mathbb{F}_p, the field of modulo p. The of a finite field is the prime p, which determines the smallest number of times the multiplicative identity must be added to itself to yield zero, and it divides the order of the field. Finite fields are typically constructed as quotient rings of the over \mathbb{F}_p by an of degree n, such as \mathbb{F}_{p^n} \cong \mathbb{F}_p / (f(x)) where f(x) is irreducible. Key structural properties include the of nonzero elements forming a of order p^n - 1, enabling the existence of primitive elements (generators) that can represent all nonzero elements as powers. The additive structure is a over \mathbb{F}_p of dimension n. Finite fields play a crucial role in numerous applications, including error-correcting codes in digital communications, cryptographic protocols such as and the , and combinatorial designs in computer science and engineering. They also underpin advanced topics in , such as the proof of via Gauss sums, and algebraic geometry over finite fields.

Definition and Properties

Definition

A finite field is a with a finite number of elements. It is an consisting of a set equipped with two binary operations, and , that satisfy the field axioms: the set forms an under (with an additive 0 and additive inverses for each element), is commutative, associative and distributive over , there is a multiplicative 1 distinct from 0, and every nonzero element has a multiplicative inverse. Finite fields are denoted by \mathbb{F}_q or GF(q), where q is the number of elements in the field, known as its . The q must be a , specifically q = p^n for some prime p and positive n \geq 1. This distinguishes finite fields from fields like the rational numbers \mathbb{Q}. The concept of finite fields originated from Évariste Galois's work on the theory of equations in 1830, where he implicitly constructed such structures while studying solvability by radicals. The term "Galois field" was coined by Leonard Eugene Dickson in his 1901 monograph Linear Groups, with an Exposition of the Galois Field Theory, which systematized their theory and introduced the notation GF(p^n). The smallest finite field is \mathbb{F}_2, with elements \{[0](/page/0), [1](/page/1)\} and operations defined modulo 2: addition is $0 + 0 = [0](/page/0), $0 + 1 = [1](/page/1), $1 + 1 = [0](/page/0); multiplication is $0 \cdot [0](/page/0) = [0](/page/0), $0 \cdot [1](/page/1) = [0](/page/0), [1](/page/1) \cdot [1](/page/1) = [1](/page/1). The additive inverse of is , of is , and the multiplicative inverse of is .

Basic properties

Every finite field F of order q has prime characteristic p and is a over its prime subfield \mathbb{F}_p of n, where q = p^n. This structure arises because the prime subfield \mathbb{F}_p embeds naturally into F, and by elements of \mathbb{F}_p is compatible with the field operations in F. The n determines the size of F as a , with \{1, \alpha, \alpha^2, \dots, \alpha^{n-1}\} forming a basis for some element \alpha \in F. As a field, F contains no zero divisors, meaning that if ab = 0 for a, b \in F, then either a = 0 or b = 0. Furthermore, every non-zero element of F has a : for any a \in F with a \neq 0, there exists b \in F such that ab = 1. This property ensures that division is possible for non-zero elements, distinguishing fields from more general rings. The non-zero elements of F form the multiplicative group F^\times of order q-1. By Lagrange's theorem applied to this finite abelian group, the order of any subgroup of F^\times divides q-1. This divisibility condition has implications for the structure of subgroups and the existence of elements of specific orders within F^\times.

Characteristic and prime subfield

In a finite field F, the characteristic is the smallest positive integer p such that p \cdot 1 = 0, where $1 denotes the multiplicative , and this p must be a . This follows from the fact that the additive order of $1 divides the order of any element in the field, and the cannot be composite without violating the field axioms. The characteristic p divides the order q of the finite field F, so q = p^n for some positive integer n \geq 1. This structure ensures that F is a vector space of dimension n over the field of p elements, though the focus here is on the foundational role of the characteristic in determining the field's arithmetic. The prime subfield of F is the smallest subfield, generated by repeated addition and multiplication of the identity element $1. Specifically, it consists of the elements \{0, 1, 2 \cdot 1, \dots, (p-1) \cdot 1\}, where addition and multiplication are performed in F, and this subfield is isomorphic to \mathbb{Z}/p\mathbb{Z}, the integers modulo p. Every finite field contains exactly one such prime subfield, which serves as the embedding of \mathbb{F}_p within F. This uniqueness arises because any subfield must contain the multiples of $1 and be closed under the field operations, leading to the of all subfields being precisely this prime subfield.

Existence and Uniqueness

Fields of prime order

A finite field of prime order p, where p is a , exists and is explicitly constructed as the \mathbb{Z}/p\mathbb{Z}, the integers p. The elements of this field, denoted \mathbb{F}_p, are the residue classes {{grok:render&&&type=render_inline_citation&&&citation_id=0&&&citation_type=wikipedia}}, {{grok:render&&&type=render_inline_citation&&&citation_id=1&&&citation_type=wikipedia}}, \dots, [p-1], with the operations of and multiplication defined by performing the usual integer operations and then reducing p. Because p is prime, \mathbb{Z}/p\mathbb{Z} is an with every nonzero element possessing a , thereby forming a . In \mathbb{F}_p, the is p, and every element a satisfies a^p = a. This equation holds for all a \in \mathbb{F}_p as a direct consequence of applied in the integers modulo p. Consequently, the x^p - x has every element of \mathbb{F}_p as a root. The x^p - x factors completely over \mathbb{F}_p as x^p - x = \prod_{a \in \mathbb{F}_p} (x - a), with all roots distinct, making \mathbb{F}_p the of x^p - x over itself. This factorization arises from the in characteristic p, where (x + y)^p = x^p + y^p, allowing the polynomial to be expressed as a product of linear terms corresponding to the field elements. Since the order p is prime, \mathbb{F}_p admits no proper subfields: any subfield would have order dividing p, so either 1 or p, but a field cannot have exactly one element. Up to isomorphism, there is a unique field of order p.

Uniqueness theorem

The uniqueness theorem for finite fields asserts that for every prime power q = p^n, where p is a prime and n is a positive integer, there exists exactly one field of order q up to isomorphism. To establish this, first note that in any F of q, every element \alpha \in F satisfies the equation \alpha^q = \alpha. This follows because the F^\times has q-1, so by , \alpha^{q-1} = 1 for \alpha \neq 0, implying \alpha^q = \alpha; the case \alpha = 0 holds trivially. Thus, F is contained in the of the x^q - x over the prime subfield \mathbb{F}_p. Moreover, x^q - x is separable over \mathbb{F}_p because its q x^{q-1} - 1 = -1 (since q = 0 in p) is a nonzero constant, ensuring no multiple . The set of of x^q - x in an of \mathbb{F}_p consists of q distinct elements and forms a subfield, as it is closed under ((\alpha + \beta)^q = \alpha^q + \beta^q = \alpha + \beta), , and additive/multiplicative inverses (the latter following from the group structure). Thus, this subfield has q and is the of x^q - x over \mathbb{F}_p, proving the existence of a of q. Any two such splitting fields are isomorphic over \mathbb{F}_p by the of splitting fields for separable polynomials. Therefore, any two finite fields of q are isomorphic as extensions of \mathbb{F}_p, and hence as fields. The multiplicative groups of both fields are cyclic of q-1, providing an additional structural confirmation, though the argument suffices for .

Constructions

Prime fields

The prime fields, also known as fields of prime , are the simplest finite fields and serve as the foundational building blocks for more complex constructions. For a p, the prime field \mathbb{F}_p is constructed as the \mathbb{Z}/p\mathbb{Z}, consisting of the residue classes of the integers p, represented by the elements \{0, 1, 2, \dots, p-1\}. This is equipped with the standard operations of and inherited from the integers, performed p: for a, b \in \mathbb{Z}, the is (a + b) \mod p and the product is (a \cdot b) \mod p. Since p is prime, \mathbb{Z}/p\mathbb{Z} has no zero divisors, meaning that if ab \equiv 0 \pmod{p} for a, b \in \mathbb{Z}/p\mathbb{Z}, then either a \equiv 0 \pmod{p} or b \equiv 0 \pmod{p}. This property, combined with the existence of additive and multiplicative inverses for every nonzero element (by , as \gcd(k, p) = 1 for $1 \leq k < p), ensures that \mathbb{F}_p is indeed a field. The additive inverse of a is -a \mod p, and the multiplicative inverse of a nonzero a is the unique b such that ab \equiv 1 \pmod{p}. A concrete example is the field \mathbb{F}_3 = \{0, 1, 2\}, where addition and multiplication are modulo 3. Here, $2 + 1 \equiv 0 \pmod{3} and $2 \cdot 2 \equiv 1 \pmod{3}, illustrating how the operations wrap around the prime modulus to form a closed algebraic structure. Up to isomorphism, \mathbb{F}_p is the unique field of order p.

Polynomial extensions

Finite fields of order p^n for a prime p and integer n > 1 are constructed as quotient rings of the over the prime . Specifically, \mathbb{F}_{p^n} is isomorphic to \mathbb{F}_p / (f(x)), where f(x) is a monic of degree n over \mathbb{F}_p. This construction leverages the fact that the ideal generated by an irreducible polynomial is maximal in the , ensuring the quotient is a . The elements of \mathbb{F}_{p^n} are represented as residue classes of polynomials in \mathbb{F}_p with degree strictly less than n, typically written in the form a_0 + a_1 \alpha + \cdots + a_{n-1} \alpha^{n-1}, where each a_i \in \mathbb{F}_p and \alpha is the image of x in the quotient (a root of f(x)). There are precisely p^n such elements, as each of the n coefficients can take p possible values. Addition of two elements is performed by adding their corresponding polynomials coefficient-wise, with coefficients reduced modulo p. Multiplication proceeds by first multiplying the representing in the usual manner and then reducing the result f(x) to obtain a of less than n. This step uses the relation f(\alpha) = 0, allowing powers of \alpha of n or higher to be expressed as linear combinations of lower powers via the coefficients of f(x). For example, if f(x) = x^n + c_{n-1} x^{n-1} + \cdots + c_0, then \alpha^n = -c_{n-1} \alpha^{n-1} - \cdots - c_0. The isomorphism \mathbb{F}_{p^n} \cong \mathbb{F}_p / (f(x)) holds for any choice of monic irreducible f(x) of degree n, and all such quotients are isomorphic to each other as fields. Thus, the structure of \mathbb{F}_{p^n} is independent of the specific irreducible polynomial selected for the construction.

Small finite field examples

One of the smallest non-prime finite fields is \mathbb{F}_4, constructed as the \mathbb{F}_2/(p(x)) where p(x) = x^2 + x + 1 is the unique monic of degree 2 over \mathbb{F}_2. Let \omega denote a root of p(x), satisfying \omega^2 = \omega + 1. The elements of \mathbb{F}_4 are then \{0, [1](/page/1), \omega, \omega + 1\}, forming a 2-dimensional over \mathbb{F}_2 with basis \{[1](/page/1), \omega\}. Addition in \mathbb{F}_4 is componentwise modulo 2, while uses the \omega^2 = \omega + [1](/page/1) and reduces higher powers accordingly. The following tables illustrate these operations: Addition table:
+\omega\omega + [1](/page/1)
\omega\omega + [1](/page/1)
\omega + [1](/page/1)\omega
\omega\omega\omega + [1](/page/1)
\omega + [1](/page/1)\omega + [1](/page/1)\omega
Multiplication table:
\cdot01\omega\omega + 1
00000
101\omega\omega + 1
\omega0\omega\omega + 11
\omega + 10\omega + 11\omega
These tables confirm that \mathbb{F}_4 satisfies the field axioms, with every nonzero element having a multiplicative inverse. The finite field \mathbb{F}_8 arises as \mathbb{F}_2/(q(x)) with q(x) = x^3 + x + 1, an of degree 3 over \mathbb{F}_2. Let \alpha be a of q(x), so \alpha^3 = \alpha + 1; this \alpha generates the and serves as a primitive element, meaning its powers \alpha^0, \alpha^1, \dots, \alpha^6 yield all nonzero elements. The elements are in \alpha of degree less than 3 with coefficients in \mathbb{F}_2: \{0, 1, \alpha, \alpha + 1, \alpha^2, \alpha^2 + 1, \alpha^2 + \alpha, \alpha^2 + \alpha + 1\}. Operations follow polynomial arithmetic q(x). For \mathbb{F}_9, consider the extension \mathbb{F}_3/(r(x)) where r(x) = x^2 + 1 is irreducible over \mathbb{F}_3, as it has no (neither 0, 1, nor 2 satisfies x^2 \equiv -1 \pmod{3}, and -1 \equiv 2 \pmod{3}). Let i be a , so i^2 = -1 = 2 in \mathbb{F}_3. The elements are a + b i for a, b \in \{0, 1, 2\}, with addition and multiplication performed modulo 3 using the relation i^2 = 2. This representation highlights \mathbb{F}_9 as adjoining a square of 2 to \mathbb{F}_3. The field \mathbb{F}_{16} is obtained via \mathbb{F}_2/(s(x)) with s(x) = x^4 + x + 1, a monic of degree 4 over \mathbb{F}_2 that is also , generating the with a \beta satisfying \beta^4 = \beta + 1. Elements are polynomials in \beta of degree less than 4 over \mathbb{F}_2, totaling , with arithmetic modulo s(x). This construction is commonly used for its efficiency in applications requiring a element.

Multiplicative Structure

Cyclic multiplicative group

The multiplicative group \mathbb{F}_q^\times of a finite field \mathbb{F}_q consists of the nonzero elements under multiplication and has order q-1. This group is cyclic. To prove cyclicity, it suffices to show the existence of an element of order d for every positive divisor d of q-1. Let N(d) denote the number of elements g \in \mathbb{F}_q^\times satisfying g^d = 1, i.e., those whose order divides d. The equation x^d - 1 = 0 is a polynomial equation of degree d over the field \mathbb{F}_q, so it has at most d roots. However, since d \mid q-1, the polynomial x^d - 1 divides x^{q-1} - 1, and the latter equation x^{q-1} - 1 = 0 has exactly q-1 distinct roots (all nonzero elements of \mathbb{F}_q). Moreover, the characteristic p of \mathbb{F}_q does not divide q-1 (hence does not divide d), so x^d - 1 is separable: its derivative d x^{d-1} shares no common roots with it, as d \not\equiv 0 \pmod{p} and roots are nonzero. Thus, x^d - 1 splits into d distinct linear factors over \mathbb{F}_q, yielding exactly N(d) = d solutions. Let \psi(d) be the number of elements of exact order d. Then, N(d) = \sum_{e \mid d} \psi(e) for each d \mid q-1. By Möbius inversion over the poset of divisors, \psi(d) = \sum_{e \mid d} \mu(d/e) \, N(e), where \mu is the . Substituting N(e) = e gives \psi(d) = \sum_{e \mid d} \mu(d/e) \, e = \phi(d), Euler's totient function. Since \phi(d) > 0 for all d \geq 1, there are exactly \phi(d) elements of order d, and in particular \phi(q-1) > 0 elements of order q-1, generating the full group. A of \mathbb{F}_q^\times is called a of \mathbb{F}_q, and their number is thus \phi(q-1). For example, in \mathbb{F}_5^\times = \{1, 2, 3, 4\} of order 4, the 2 has powers $2^1 = 2, $2^2 = 4, $2^3 = 3, $2^4 = 1 (modulo 5), generating the group and confirming it is ; there are \phi(4) = 2 such elements (2 and 3).

Primitive elements

In a finite field \mathbb{F}_q of order q = p^n where p is prime and n \geq 1, a is a g \in \mathbb{F}_q^\times of the cyclic multiplicative group \mathbb{F}_q^\times, which has order q-1. The powers of such a g thus exhaust all nonzero elements: \{g^0, g^1, \dots, g^{q-2}\} = \mathbb{F}_q \setminus \{0\}. The existence of elements follows from the cyclicity of \mathbb{F}_q^\times. The number of primitive elements in \mathbb{F}_q is exactly \phi(q-1), where \phi denotes ; hence, their proportion among the nonzero elements is \phi(q-1)/(q-1). The minimal polynomial of a primitive element over the prime subfield \mathbb{F}_p is an irreducible of degree n. To determine whether a given element g \in \mathbb{F}_q^\times is primitive, compute its , which must equal q-1. Assuming the prime factorization of q-1 = \prod p_i^{e_i} is known, this requires verifying g^{(q-1)/p_i} \neq 1 for each distinct prime p_i of q-1, as failure for any p_i would imply a proper of q-1 as the . For example, in the prime field \mathbb{F}_7, the element $3 is primitive because its powers cycle through all nonzero elements: $3^1 \equiv 3, $3^2 \equiv 2, $3^3 \equiv 6, $3^4 \equiv 4, $3^5 \equiv 5, $3^6 \equiv 1 \pmod{7}. To confirm using the test (with q-1=6=2 \cdot 3), check $3^{6/2} = 3^3 \equiv 6 \not\equiv 1 and $3^{6/3} = 3^2 \equiv 2 \not\equiv 1 \pmod{7}.

Roots of unity and subgroups

The \mathbb{F}_q^\times of a finite field \mathbb{F}_q is cyclic of q-1. Consequently, for each positive d of q-1, there exists a of \mathbb{F}_q^\times of d. This structure arises because cyclic groups contain precisely one for each of their , and the cyclicity of \mathbb{F}_q^\times ensures no additional subgroups exist. The m-th roots of unity in \mathbb{F}_q are the solutions to the equation x^m - 1 = 0. These elements exist and number exactly m if and only if m divides q-1, in which case they form a cyclic of \mathbb{F}_q^\times of m. This is the of the x \mapsto x^m from \mathbb{F}_q^\times to itself, which has degree m precisely when m \mid q-1. The primitive m-th roots of unity are the roots of the m-th \Phi_m(x), which has degree \phi(m) over \mathbb{Q}, where \phi is . Over \mathbb{F}_q with \gcd(q, m) = 1, \Phi_m(x) factors into \phi(m)/\ord_m(q) distinct irreducible polynomials, each of degree \ord_m(q), where \ord_m(q) denotes the multiplicative order of q modulo m. This factorization reflects the Galois structure of the extension generated by a primitive m-th , whose degree over \mathbb{F}_q is \ord_m(q). A concrete example occurs in \mathbb{F}_{13}, where q-1 = 12 and the quadratic residues (nonzero squares) form the unique of 2 in \mathbb{F}_{13}^\times, hence of order 6. These residues are $1, 3, 4, 9, 10, 12 modulo 13, comprising half the nonzero elements as expected for the image of the squaring map in . This corresponds to the 2nd roots of unity extended by higher powers, illustrating the divisor-based structure.

Automorphisms

Frobenius endomorphism

In a finite field \mathbb{F}_q of p, where q = p^n for some prime p and positive integer n, the \phi: \mathbb{F}_q \to \mathbb{F}_q is defined by \phi(a) = a^p for all a \in \mathbb{F}_q. This map is a field homomorphism because, in p, the binomial theorem implies (a + b)^p = a^p + b^p, so \phi(a + b) = \phi(a) + \phi(b), and \phi(ab) = (ab)^p = a^p b^p = \phi(a) \phi(b). Since \mathbb{F}_q is finite, \phi is injective (its kernel is \{0\}) and hence bijective, making it an automorphism of the field. The powers of \phi satisfy \phi^k(a) = a^{p^k} for each k, and in particular \phi^n(a) = a^{p^n} = a^q = a for all a \in \mathbb{F}_q, since every element of \mathbb{F}_q is a root of the polynomial x^q - x. Thus, \phi^n is the identity map, so the order of \phi divides n; in fact, the order is exactly n. The fixed field of \phi, consisting of elements a such that \phi(a) = a (i.e., a^p = a), is precisely the prime subfield \mathbb{F}_p. For any \alpha \in \mathbb{F}_{p^n}, the minimal polynomial of \alpha over \mathbb{F}_p divides x^{p^n} - x, and the roots of this minimal polynomial—the conjugates of \alpha under the action of the —are exactly \alpha, \alpha^p, \dots, \alpha^{p^{d-1}}, where d is the degree of the minimal polynomial. The \mathrm{Gal}(\mathbb{F}_{p^n}/\mathbb{F}_p) is cyclic of order n and generated by \phi.

Galois group structure

The Galois group of the extension \F_{p^n}/\F_p, denoted \Gal(\F_{p^n}/\F_p), is cyclic of order n. This group is generated by the Frobenius automorphism \phi: a \mapsto a^p. The elements of \Gal(\F_{p^n}/\F_p) are precisely the automorphisms \sigma_k for k = 0, 1, \dots, n-1, defined by \sigma_k(a) = a^{p^k} for all a \in \F_{p^n}. These automorphisms are distinct and form a basis for the group structure, with \sigma_k^n = \id for each k. The relation \sigma_k = \phi^k highlights the cyclic nature generated by the Frobenius map. For subextensions \F_{p^m}/\F_p where m divides n, the Galois group \Gal(\F_{p^n}/\F_{p^m}) is cyclic of order n/m, serving as a subgroup of \Gal(\F_{p^n}/\F_p) of index m. The fundamental theorem of Galois theory establishes a bijection between the intermediate fields of \F_{p^n}/\F_p and the subgroups of \Gal(\F_{p^n}/\F_p); specifically, the fixed field of the subgroup generated by \phi^m is \F_{p^m}. This correspondence underscores the lattice of subfields aligning with the divisor lattice of n. As an example, consider the extension \F_{p^4}/\F_p. The is cyclic of 4, with subgroups of orders 1, 2, and 4 corresponding to the subfields \F_p, \F_{p^2}, and \F_{p^4}, respectively. The quadratic subfield \F_{p^2} is fixed by the subgroup \langle \phi^2 \rangle of 2. For a case with p=2 and n=2, the extension \F_4/\F_2 has of 2 generated by \phi_2: x \mapsto x^2, where the nontrivial swaps the elements, such as mapping a of x^2 + x + 1 = 0 to its conjugate.

Polynomial Factorization

Distinct degree factorization

In finite fields \mathbb{F}_q, every non-constant polynomial f(x) \in \mathbb{F}_q factors uniquely into a product of irreducible polynomials, and the distinct-degree factorization groups these irreducible factors by their degrees, providing an initial decomposition f = \prod_d g_d^{e_d} where each g_d is the square-free product of all distinct irreducible factors of degree d and e_d is related to the multiplicities, though for square-free f, the exponents are 1. This approach leverages the structure of extension fields and the Frobenius automorphism to isolate factors of specific degrees without finding the individual irreducibles. The key tool is the polynomial x^{q^d} - x, which factors as the product of all monic irreducible over \mathbb{F}_q whose divide d. This follows from the fact that the roots of x^{q^d} - x are precisely the elements of the extension \mathbb{F}_{q^d}, and by unique factorization in \mathbb{F}_q, it is the product over all irreducibles whose roots lie in subfields of dividing d. The \sigma: \alpha \mapsto \alpha^q generates the of \mathbb{F}_{q^d}/\mathbb{F}_q, and an element \alpha \in \mathbb{F}_{q^d} has minimal polynomial dividing d \sigma^d(\alpha) = \alpha, meaning \alpha^{q^d} = \alpha, so x^{q^d} - x vanishes on such elements. For a square-free polynomial f \in \mathbb{F}_q of degree n, the distinct-degree factorization writes f = \prod_{d=1}^n g_d, where g_d is the product of all distinct irreducible factors of f of exact degree d. The algorithm computes this iteratively using greatest common divisors: set V_0 = 1; for d = 1 to n, compute V_d = \gcd(f, x^{q^d} - x), then g_d = V_d / V_{d-1}. Each irreducible factor \phi of degree d divides x^{q^d} - x (since its roots satisfy the equation in \mathbb{F}_{q^d}) but not x^{q^k} - x for proper divisors k < d of the degree (since the roots generate an extension of exact degree d), ensuring the isolation of exact-degree components. Computing the gcds requires efficient modular exponentiation for x^{q^d} \mod f, which can be done in polynomial time using repeated squaring. For general (possibly non-square-free) f, first decompose into square-free parts using \gcd(f, f') and higher derivatives, then apply the distinct-degree step to each square-free component, with multiplicities determined from the decomposition. As an example, consider factoring f(x) = x^6 + x^3 + 1 over \mathbb{F}_2 (where q=2), which turns out to be irreducible of degree 6. For d=1, compute \gcd(f, x^2 + x): since f(0) = 1 \neq 0 and f(1) = 1 + 1 + 1 = 1 \neq 0, the gcd is 1, so no degree-1 factors. For d=2, compute \gcd(f, x^4 + x): using the Euclidean algorithm, divide f by x^4 + x to get remainder 1, so gcd is 1, indicating no degree-2 factors. For d=3, compute \gcd(f, x^8 + x): reducing x^8 + x \mod f yields x^5 + x^2 + x, and continuing the Euclidean algorithm gives gcd 1, as expected since the roots do not lie in \mathbb{F}_8. The process continues to d=6, where \gcd(f, x^{64} + x) = f, confirming g_6 = f and that all irreducible factors have degree 6. This example illustrates how the method progressively eliminates lower-degree possibilities using .

Counting irreducible polynomials

The number of monic irreducible polynomials of degree n over the finite field \mathbb{F}_q, denoted I_q(n), is given by the formula I_q(n) = \frac{1}{n} \sum_{d \mid n} \mu(d) q^{n/d}, where \mu is the Möbius function. This formula arises from Möbius inversion applied to the structure of finite field extensions. The field \mathbb{F}_{q^n} has exactly q^n elements, and each element lies in a unique minimal polynomial over \mathbb{F}_q, which is monic and irreducible of some degree d dividing n. Each such irreducible polynomial of degree d has precisely d roots in \mathbb{F}_{q^n}. Thus, q^n = \sum_{d \mid n} d \, I_q(d). Applying Möbius inversion to this relation yields the expression for I_q(n). For large n, the term corresponding to d=1 dominates the sum, since the other summands involve smaller exponents of q. Consequently, I_q(n) \sim \frac{q^n}{n}. For example, over \mathbb{F}_2, the divisors of 3 are 1 and 3, so I_2(3) = \frac{1}{3} \left( \mu(1) \cdot 2^3 + \mu(3) \cdot 2^1 \right) = \frac{1}{3} (8 - 2) = 2. The two monic irreducible polynomials of degree 3 are x^3 + x + 1 and x^3 + x^2 + 1.

Berlekamp's algorithm overview

Berlekamp's algorithm, introduced in 1967, is a deterministic method for factoring square-free polynomials over finite fields into irreducible factors, building on prior steps such as distinct-degree factorization to handle equal-degree components. The approach leverages linear algebra over the finite field \mathbb{F}_q to identify and separate the irreducible constituents of a given polynomial. For a square-free monic polynomial h(x) of degree kd over \mathbb{F}_q, where k is the number of irreducible factors each of degree d, the algorithm constructs the Berlekamp subalgebra consisting of residue classes of polynomials v(x) modulo h(x) satisfying v(x)^q \equiv v(x) \pmod{h(x)}. This condition defines the vectors in the null space of the linear transformation Q - I, where I is the identity and Q acts on a polynomial v(x) evaluated at a root \alpha of h(x) by Q(v)(\alpha) = v(\alpha^q). The dimension of this null space equals k, the number of irreducible factors. The algorithm proceeds in several key steps. First, the matrix representation of Q is computed relative to the standard monomial basis \{1, x, \dots, x^{kd-1}\}, often via repeated computations to evaluate powers like x^{q^i} \mod h(x). Next, the kernel of Q - I is found using over \mathbb{F}_q, yielding a basis for the Berlekamp subalgebra. To split h(x), elements c \in \mathbb{F}_q are tested: for each basis vector v in the kernel (beyond the trivial constants), compute \gcd(h(x), v(x) - c); non-trivial factors arise when c is not an eigenvalue corresponding to all factors simultaneously. This process recursively factors the polynomial until irreducibles are obtained. The time complexity of Berlekamp's algorithm is O(d^3) arithmetic operations in \mathbb{F}_q for a polynomial of degree d, dominated by the linear algebra over a d \times d matrix.

Applications

Coding theory

Finite fields form the foundation for , a prominent class of error-correcting codes in coding theory. These codes are constructed as evaluation codes over a finite field \mathbb{F}_q: a message corresponds to a polynomial f(x) of degree less than k, which is evaluated at n distinct points \alpha_1, \dots, \alpha_n \in \mathbb{F}_q to produce the codeword (f(\alpha_1), \dots, f(\alpha_n)), with code length n \leq q. This algebraic structure leverages the properties of polynomials over finite fields to ensure reliable error detection and correction in data transmission and storage. The minimum distance of a Reed-Solomon code is d = n - k + 1, achieving equality in the Singleton bound and classifying it as a maximum distance separable (MDS) code. To correct up to t errors, the generator polynomial g(x) is chosen to have degree $2t = n - k, typically as the least common multiple of minimal polynomials corresponding to $2t consecutive powers of a primitive element. This design allows the code to correct any t symbol errors or detect up to $2t errors, with the error-correcting capability bounded by t = \lfloor (d-1)/2 \rfloor. Reed-Solomon codes over \mathbb{F}_{256} are widely applied in practical systems, such as QR codes for robust 2D barcode reading and DVDs for correcting scratches and defects during optical data retrieval. Finite fields enable algebraic decoding methods, where syndromes computed in \mathbb{F}_q from received symbols quantify errors and guide correction without exhaustive search. Polynomial factorization over finite fields supports this process by solving for the error locator polynomial.

Cryptography

Finite fields play a central role in several cryptographic s, particularly those relying on the hardness of computational problems within their multiplicative groups or on elliptic curves defined over them. The discrete logarithm problem (DLP) in the multiplicative group \mathbb{F}_p^* of a prime field involves finding an integer x such that g^x \equiv h \pmod{p}, where g is a generator and h \in \mathbb{F}_p^*. This problem underpins the security of and encryption schemes, as the best general algorithms, including variants of the index calculus method, solve it in subexponential time complexity L_p(1/2, c) for some constant c > 0, where L_p(a, b) = \exp((b + o(1)) (\log p)^a (\log \log p)^{1-a}). The Diffie-Hellman uses this hardness to enable secure shared key agreement over an insecure channel without prior secrets, by having parties compute g^{ab} \pmod{p} from public values g^a \pmod{p} and g^b \pmod{p}. Similarly, the ElGamal provides public-key based on the DLP, where consists of pairs (g^k \pmod{p}, m \cdot y^k \pmod{p}) with public key y = g^x \pmod{p} and message m, decryptable only by the private key x. Elliptic curve cryptography (ECC) extends these ideas by defining elliptic curves over finite fields \mathbb{F}_p or extensions \mathbb{F}_{p^n}, typically in Weierstrass form y^2 = x^3 + a x + b with a, b \in \mathbb{F}_p satisfying the non-singularity condition $4a^3 + 27b^2 \neq 0. The points on the curve, including the point at infinity, form an abelian group under a chord-and-tangent addition law, enabling analogs of DLP-based protocols with smaller key sizes for equivalent security due to the elliptic curve discrete logarithm problem (ECDLP) being harder to solve than the classical DLP. ECC was proposed independently by Neal Koblitz and Victor Miller in 1985, with applications including the Elliptic Curve Diffie-Hellman (ECDH) key exchange and Elliptic Curve Digital Signature Algorithm (ECDSA). The (), standardized as Rijndael, incorporates finite fields in its byte substitution step via the field \mathbb{F}_{2^8} = \mathbb{F}_2/(x^8 + x^4 + x^3 + x + 1). The is constructed by taking the in \mathbb{F}_{2^8} (with 0 mapping to itself) followed by an over \mathbb{F}_2, represented by the matrix multiplication and addition that ensures non-linearity and resistance to . This design provides and properties essential for AES's security as a symmetric . Pairing-based cryptography leverages bilinear maps, such as the Ate pairing, defined on elliptic curves over finite fields with small embedding degrees k, where the pairing e: E(\mathbb{F}_q) \times E(\mathbb{F}_{q^k}) \to \mathbb{G}_T satisfies bilinearity, non-degeneracy, and efficient computability. The Ate pairing, introduced by Hess, Smart, and Vercauteren, shortens the Miller loop compared to the Tate pairing, enabling applications like identity-based encryption and short signatures on pairing-friendly curves such as Barreto-Naehrig curves. Its security relies on the hardness of problems like the bilinear Diffie-Hellman in these finite field settings.

Combinatorial designs

Finite fields provide a foundational framework for constructing finite geometries, which serve as key examples of combinatorial designs known as block designs. The affine of order q, denoted AG(2,q), is constructed over the finite field \mathbb{F}_q as the two-dimensional \mathbb{F}_q^2, consisting of q^2 points. Lines in AG(2,q) are the cosets of one-dimensional subspaces, partitioned into q+1 parallel classes, each containing q parallel lines with q points per line, yielding a total of q(q+1) lines. Every pair of distinct points determines a unique line, and every pair of distinct lines either intersect at exactly one point or are parallel, satisfying the axioms of a $2-( q^2, q, 1) design. The of order q, denoted PG(2,q), extends this and is derived from the three-dimensional \mathbb{F}_q^3, where points are the one-dimensional subspaces (with (q^3 - 1)/(q - 1) = q^2 + q + 1 points) and lines are the two-dimensional subspaces (also q^2 + q + 1 lines). Each line contains q + 1 points, each point lies on q + 1 lines, and any two points or lines intersect at exactly one point, forming a symmetric $2-( q^2 + q + 1, q + 1, 1) . This arises from the over \mathbb{F}_q, enabling systematic enumeration and incidence relations. A prominent application involves difference sets, subsets D of a group G of v such that every non-identity element appears exactly \lambda times as a difference d_1 d_2^{-1} for d_1, d_2 \in D with d_1 \neq d_2, and |D| = k. In PG(2,q), the points can be coordinatized using the of q^2 + q + 1, where the lines correspond to Singer difference sets with parameters (v, k, \lambda) = (q^2 + q + 1, q + 1, 1). These sets, introduced by Singer in , generate symmetric designs isomorphic to the and exist whenever a finite field of q does. An illustrative example is the Paley design, constructed using quadratic residues in the finite field \mathbb{F}_p where p is a prime congruent to 3 4. The nonzero quadratic residues form the blocks of a symmetric $2-( p, (p-1)/2, (p-5)/4) design, leveraging the multiplicative character of the to ensure balanced differences. This construction, due to Paley in , yields a conference matrix that underlies various block designs. These finite field-derived designs have significant applications, including the construction of Hadamard matrices. For instance, the Paley construction produces a Hadamard matrix of order q + 1 from the Jacobian of the quadratic residues in \mathbb{F}_q when q \equiv 3 \pmod{4}, providing explicit examples that meet the Hadamard conjecture for certain orders. Additionally, such designs facilitate tight bounds in coding theory; for example, the parameters of projective plane designs imply upper limits on code sizes via the Johnson bound, establishing fundamental constraints on error-correcting codes over finite fields.

Advanced Topics

Wedderburn's little theorem

Wedderburn's little theorem, proved by Maclagan Wedderburn in 1905, states that every finite is commutative and thus a field. This result establishes that finite division rings, also known as skew fields, coincide precisely with the finite fields, as no non-commutative examples exist in the finite case. The theorem provides a sharp contrast to the infinite setting, where non-commutative division rings like Hamilton's quaternions (discovered in 1843) demonstrate that commutativity is not forced by the division ring structure alone. Wedderburn's original proof builds on his contemporaneous work in structure theory for finite algebras over fields. The center Z of a finite division ring D is shown to be a finite field \mathbb{F}_q for some q, as it is a finite commutative . Then, D is a finite-dimensional over Z, say of n \geq 1, so |D| = q^n. To establish commutativity, Wedderburn analyzes the orders of elements in the D^\times, which has order q^n - 1, using number-theoretic properties of these orders and the structure of finite linear groups to derive a unless n = 1, in which case D = Z is commutative. A streamlined modern proof, due to Ernst Witt in 1931, employs elementary via the class equation for D^\times and basic facts about cyclotomic polynomials over \mathbb{F}_q. After establishing the center as \mathbb{F}_q and dimension n, the proof shows that the equation x^{q-1} = 1 has exactly q-1 solutions in D (the nonzero elements of Z), by arguing that any solution generates a commutative sub-division ring isomorphic to a subfield of \mathbb{F}_q and must centralize all of D via properties of irreducible representations or polynomial degrees. Assuming n > 1 leads to more solutions than the degree q-1 allows without violating the finite dimensionality or the irreducibility of cyclotomic factors, yielding the contradiction and forcing n = 1.

Algebraic closure

The algebraic closure of a finite field \mathbb{F}_q, denoted \overline{\mathbb{F}_q}, is constructed as the direct limit (union) of all finite extensions \mathbb{F}_{q^n} for n = 1, 2, \dots. This field is of \mathbb{F}_q in which every non-constant with coefficients in \mathbb{F}_q splits completely into linear factors. By definition, \overline{\mathbb{F}_q} is algebraically closed, meaning it contains all roots of such polynomials, and every element of \overline{\mathbb{F}_q} is algebraic over \mathbb{F}_q, with no transcendental elements. As a union of finite fields, \overline{\mathbb{F}_q} is infinite yet countable, since there are countably many finite extensions and each \mathbb{F}_{q^n} is finite. Every element \alpha \in \overline{\mathbb{F}_q} lies in some finite extension \mathbb{F}_{q^k} and thus satisfies the equation x^{q^k} - x = 0, reflecting the fact that the elements of \overline{\mathbb{F}_q} are precisely those algebraic over \mathbb{F}_q. The \mathrm{Gal}(\overline{\mathbb{F}_q}/\mathbb{F}_q) is isomorphic to the profinite \hat{\mathbb{Z}} of the integers under , endowed with the profinite . This group is topologically generated by the Frobenius automorphism \phi: x \mapsto x^q, which has infinite order and whose powers correspond to the action on the finite extensions.

Function fields over finite fields

Function fields over finite fields serve as geometric analogs to number fields in arithmetic , where the role of the rational numbers \mathbb{Q} is played by the rational function field over a finite . The rational function field \mathbb{F}_q(x) is defined as the quotient field of the \mathbb{F}_q, consisting of all ratios of polynomials in \mathbb{F}_q with nonzero denominator, where \mathbb{F}_q is the finite with q elements. This has transcendence degree 1 over \mathbb{F}_q and provides a foundational example for studying more general function fields, which are finite extensions of \mathbb{F}_q(x). These extensions correspond to algebraic s over \mathbb{F}_q, capturing the of the curve through the 's . More generally, the function field of a smooth projective C of g over \mathbb{F}_q, denoted k(C), is the field of rational functions on C, which is a finite of \mathbb{F}_q(x) of degree 2 if C is hyperelliptic (e.g., defined by y^2 = f(x) with \deg f = 2g+1 or $2g+2), for instance. fields arise when C is an , so g = 1, and k(C) encodes the arithmetic of points on C over extensions of \mathbb{F}_q. The g measures the complexity of the curve, analogous to the in number fields, and influences key invariants like the number of places (primes) in the field. A fundamental tool for analyzing these fields is the zeta function Z(u), defined as Z(u) = \exp\left( \sum_{n=1}^\infty \frac{\#C(\mathbb{F}_{q^n})}{n} u^n \right), where \#C(\mathbb{F}_{q^n}) is the number of \mathbb{F}_{q^n}-rational points on C. The , formulated in , posit that for a smooth projective variety over \mathbb{F}_q, the reciprocal roots of the zeta function lie on the circle of radius q^{w/2} in the complex plane (where w is the weight corresponding to the cohomology degree), and these were fully proved by Pierre Deligne in 1974 using étale cohomology. For curves, the Weil conjectures imply precise bounds on the number of rational points. A key example is the Hasse-Weil bound: for a smooth projective C of g over \mathbb{F}_q, the number of \mathbb{F}_q-rational points satisfies |\#C(\mathbb{F}_q) - (q + 1)| \leq 2g \sqrt{q}. This bound highlights the interplay between the field size q, g, and point count, with equality achieved for certain maximal curves like the Hermitian curve. In applications, Drinfeld modules provide an analog of elliptic curves within function field arithmetic: a Drinfeld \mathbb{F}_q[T]-module over a field K (extension of \mathbb{F}_q(T)) is a \phi: \mathbb{F}_q[T] \to K\{\tau\} (where \tau is the ) such that \phi_T has nonzero \tau-coefficient, enabling the study of and L-functions over function fields in a manner parallel to elliptic curves over \mathbb{Q}. Introduced by in the 1970s, these modules facilitate analogs of complex multiplication and modular forms in the function field setting.

References

  1. [1]
    Finite Field -- from Wolfram MathWorld
    A finite field is a field with a finite number of elements, also called a Galois field. Its order is always a prime or a power of a prime.Missing: authoritative sources
  2. [2]
    [PDF] notes on finite fields
    A finite field is a field with a finite underlying set. For example, Z/pZ (Fp) is a finite field where p is a prime number.Missing: authoritative sources
  3. [3]
    [PDF] Finite Fields
    1.2 Existence and uniqueness of finite fields. Now, we prove that the number of elements in a finite field must be a power of prime. Indeed, any finite field ...Missing: applications | Show results with:applications
  4. [4]
    [PDF] Introduction to finite fields - Stanford University
    Finite fields are fields with a finite number of elements, also called Galois fields. For example, Fpm has pm elements.Missing: authoritative sources
  5. [5]
    [PDF] Introduction to finite fields
    Sep 9, 2019 · A finite field is a finite set with addition and multiplication, satisfying field axioms. A simple example is Fp = {0, 1,...,p-1} with addition ...Missing: authoritative sources<|control11|><|separator|>
  6. [6]
    [PDF] Chapter 1 Finite Fields - Hyperelliptic org
    Finite fields are one of the essential building blocks in coding theory and cryptography and thus appear in many areas in IT security.
  7. [7]
    Linear groups, with an exposition of the Galois field theory
    Oct 31, 2007 · Linear groups, with an exposition of the Galois field theory. by: Dickson, Leonard E. (Leonard Eugene), 1874-. Publication date: 1901.
  8. [8]
    [PDF] Chapter 22 Finite Fields
    Let F be a finite field with characteristic field p, which must be a prime. Then Zp = {1, 2,...,p} is a subfield, and F is a finite dimensional vector space ...
  9. [9]
    [PDF] Introduction to finite fields
    Sep 16, 2013 · A finite field is a field which is, well, finite. Recall the notion of a vector space over a field. Note that an n-dimensional vector space over ...
  10. [10]
    [PDF] Finite Fields: Additive Structure
    Finite Fields: ... The prime subfield of a finite field. Claim: The ... More precisely, GF(q) may be regarded as a vector space of dimen- sion m over Fp.<|control11|><|separator|>
  11. [11]
    [PDF] how to construct them, properties of elements in a finite field, and ...
    This handout discusses how to construct finite fields, their properties, and relations. Every finite field has prime power order, and all fields of the same ...
  12. [12]
    [PDF] Basic Concept in Number Theory and Finite Fields
    a multiplicative inverse. The number of elements is always a power of a prime number. There is no GF(6) since 6 is not a power of a prime. Find remainders, ...
  13. [13]
    [PDF] 3 Finite fields and integer arithmetic - MIT Mathematics
    Feb 13, 2019 · Theorem 3.11. Every finite subgroup of the multiplicative group of a field is cyclic. Proof. Let k be a field, let G be a subgroup of k× of ...
  14. [14]
    [PDF] Finite Fields
    A finite field has a finite characteristic, the smallest positive integer n where 1+...+1^n=0. Every finite field has a characteristic of a prime number.
  15. [15]
    [PDF] 1 Characteristic and size of a finite field - People @EECS
    A finite field's characteristic is a prime number (p), and its size is |F| = pm, where m is an integer greater than or equal to 1.
  16. [16]
    [PDF] The Structure of Finite Fields - Henry Pfister
    ... multiplicative order satisfies ord(α) = q − 1. We will see shortly that such an element always exists and generates the multiplicative group. Lemma 2.10 ...
  17. [17]
    [PDF] 2 Field Theory - Dana Ernst
    The prime subfield of a field F is the subfield generated by 1F (i.e. ... Suppose F is a finite field of characteristic p. Then every element of F is ...
  18. [18]
  19. [19]
    Fermat's little theorem - PlanetMath.org
    Mar 22, 2013 · More succinctly, the group of units in a finite field is cyclic. If q is prime, then we have Fermat's little theorem.
  20. [20]
    Prime Field -- from Wolfram MathWorld
    A prime field is a finite field GF(p) for p is prime.
  21. [21]
    [PDF] Introduction to finite fields
    Sep 18, 2023 · The simplest example of a finite field is as follows. Take a prime p ∈ Z. Let Fp = Z/pZ (the quotient of the ring Z mod the ideal pZ).
  22. [22]
    [PDF] Finite Fields with Prime Power Elements
    The goal is to construct finite fields with pn elements from the polynomial rings Fp[x]. The construction will be very similar to that of Fp = Z/pZ from Z, ...
  23. [23]
    [PDF] 1. introduction to finite fields
    A finite field is a field with only finitely many elements. Exercise 1.2. Given any prime number p, show that the set Z/pZ forms a field under addition and ...
  24. [24]
    (Finite) Fields — A Primer - Math ∩ Programming
    Feb 26, 2014 · A field is a commutative ring with 0 and 1, where every non-zero element has a multiplicative inverse. A finite field has a finite number of ...Missing: authoritative sources
  25. [25]
    [PDF] Construction of Irreducible Polynomials over Finite Fields - DiVA portal
    May 22, 2014 · Abstract. In this paper we investigate some results on the construction of irreducible poly- nomials over finite fields.
  26. [26]
    [PDF] Finite fields I talked in class about the field with two elements F2 = {0 ...
    Some of these inverses exist, even for elements having no multiplicative inverse in Z: for example 3 ·10 7 = 1, so. 7 is a multiplicative inverse of 3 in Z/10Z.
  27. [27]
    Finite fields, by Rudolf Lidl and Harald Niederreiter, Second edition ...
    Finite fields are fields that have only a finite number of elements. The simplest examples are the rings of integers modulo a prime number p.
  28. [28]
    [PDF] Constructing irreducible polynomials recursively with a ... - arXiv
    Jan 23, 2023 · Let F8 = F(a) where a is a root of the monic irreducible polynomial. X3 + X + 1 ∈ F2[X]. We consider the primitive monic irreducible polynomial.Missing: F_8 F_2
  29. [29]
    [PDF] Finite Multiplicative Subgroups of a Field
    There are several ways to prove that G is cyclic. All proofs are based on the fact that the equation xd = 1 can have at most d solutions in a field F. Proof ( ...
  30. [30]
    Finite Fields - Rudolf Lidl, Harald Niederreiter - Google Books
    ... primitive element primitive polynomial Proc Prove q-polynomial quadratic character quadratic form reine angew result ring Russian S₁ sequence in F splitting ...
  31. [31]
    [PDF] Searching for Primitive Roots in Finite Fields - Victor Shoup
    By a search procedure for primitive roots in GF(pn ), we mean an algorithm that generates a subset of GF(pn ) that contains. (with high probability, in the case ...
  32. [32]
    Primitive Element - an overview | ScienceDirect Topics
    Let x = 3 be an element of the field G F 7 1 = F 7 . This element is a primitive element since it generates all the non-zero elements of F 7 . The element x = ...<|control11|><|separator|>
  33. [33]
    [PDF] Introduction to finite fields
    In an additive group, “−” is called subtraction; in a multiplicative group, “−” is called division and denoted by / or ÷.
  34. [34]
    [PDF] Finite fields: further properties
    A root α of any of these factors is a primitive (q − 1)st root of unity over Fp, and hence a primitive element of Fq. For such an α we have Fq = {0, α, α2,...,αq ...<|separator|>
  35. [35]
    Quadratic Residue -- from Wolfram MathWorld
    . For example, 4^2=6 (mod 10) , so 6 is a quadratic residue (mod 10). The entire set of quadratic residues (mod 10) are given by 1, 4, 5, 6, and 9, since. 1 ...
  36. [36]
    [PDF] Math 113, Summer 2015 Prof. Haiman Notes on finite fields 1. The ...
    The Frobenius automorphism Φ is an element of the Galois group G of F(q) over Zp. Its fixed field consists of the roots of the equation xp − x = 0 in F(q). But ...
  37. [37]
    [PDF] 1. constructing finite fields - Galois theory lecture summary
    May 14, 2018 · Well, since Gal(F/Fp) is a group, any power of φp is also an element of the Galois group. What are some other nontrivial elements of the Galois ...
  38. [38]
    [PDF] Factoring Polynomials Over Finite Fields: A Survey
    In this survey, we discuss several algorithms for the factorization of univariate polynomials over finite fields. Fq denotes a finite field with q elements. We ...
  39. [39]
    [PDF] Factoring Polynomials over Finite Fields: Asymptotic Complexity vs ...
    If we let × and logp tend to infinity at the same rate, then Berlekamp's algorithm uses 0 ( Чж ) operations, Cantor / Zassenhaus uses 0 ( × 3 log × loglog ×),.<|control11|><|separator|>
  40. [40]
    [PDF] Polynomial Codes Over Certain Finite Fields
    REED AND G. SOLOMON: Introduction. A code is a mapping from a vector space of dimension m over a finite field K (denoted by Vfl%(K)) into a vector space of ...
  41. [41]
    [PDF] Reed-Solomon Codes - University at Buffalo
    May 8, 2015 · Reed-Solomon codes have been studied a lot in coding theory. These codes are optimal in the sense that they meet the Singleton bound. (Theorem ...
  42. [42]
    [PDF] Reed Solomon Codes
    Denoted as GF(n), where n is the number of elements in the field. • A set of numbers with arithmetic operations add, subtract, multiply,.
  43. [43]
    [PDF] The Discrete Logarithm Problem - KEVIN S. McCURLEY
    This paper is a survey of the discrete logarithm problem, including the current state of algorithms for solving it, complexity issues related to the discrete ...Missing: ElGamal | Show results with:ElGamal
  44. [44]
    [PDF] Discrete logarithms in finite fields and their cryptographic significance
    Given a primitive element g of a finite field GF(q), the discrete logarithm of a nonzero element u ∈ GF(q) is that integer k, 1 ≤ k ≤ q −1, for which u = gk ...Missing: seminal | Show results with:seminal
  45. [45]
    [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 ...
  46. [46]
    [PDF] A public key cryptosystem and a signature scheme based on ...
    The paper described a public key cryptosystem and a signature scheme based on the difficulty of computing discrete logarithms over finite fields. The ...
  47. [47]
    [PDF] Elliptic Curve Cryptosystems - Evervault
    Abstract. We discuss analogs based on elliptic curves over finite fields of public key cryptosystems which use the multiplicative group of a finite field.
  48. [48]
    [PDF] Use of Elliptic Curves in Cryptography - Victor S. Miller - Evervault
    We discuss the use of elliptic curves in cryptography. In particular, we propose an analogue of the. Diffie-Hellmann key exchange protocol which appears to be ...
  49. [49]
    [PDF] FIPS 197, Advanced Encryption Standard (AES)
    Nov 26, 2001 · In matrix form, the affine transformation element of the S-box can be expressed as: 1 The various transformations (e.g., SubBytes(), ShiftRows() ...
  50. [50]
    [PDF] Projective and affine planes 1 Projective planes
    The net corresponding to the family of q−1 MOLS of order q constructed by the “finite field method” is the affine plane AG(2,q). For prime orders, no other ...Missing: source | Show results with:source
  51. [51]
    [PDF] 2 Projective planes
    1 q q 1 points; so PG 2 q is a projective plane of order q. In the infinite case, the claim follows by simple cardinal arithmetic.) 19. Page 2. 20. 2.
  52. [52]
    [PDF] Singer difference sets and the projective norm graph - arXiv
    Aug 15, 2019 · Abstract. We demonstrate a close connection between the classic planar Singer difference sets and certain norm equation systems arising from ...<|control11|><|separator|>
  53. [53]
    [PDF] Paley and the Paley graphs - arXiv
    Jan 31, 2017 · In [67] Paley gave several constructions of Hadamard matrices, based on the combinatorial properties of quadratic residues such as equation (1).
  54. [54]
    [PDF] Difference Sets From Projective Planes
    a couple of famous unsolved conjectures on planar difference sets: • Any finite projective plane admitting a Singer group is desarguesian. • Any abelian ...
  55. [55]
    A Theorem on Finite Algebras - jstor
    A THEOREM ON FINITE ALGEBRAS*. BY. J. H. MACLAGAN-WVEDDERBURN t. FPROBENIUS I and C. S. PEIRCE ? have shown that, in the domain of all real numbers, the only ...
  56. [56]
    [PDF] Wedderburn's Theorem on Division Rings
    Wedderburn's Theorem on Division Rings: A finite division ring is a field. Necessary facts: (1) If V is a vector space of dimension n over a finite field F ...Missing: little paper
  57. [57]
    [PDF] Quaternion Algebras and the Algebraic Legacy of Hamilton's ...
    Wedderburn's Theorem has proved to be a surprisingly powerful tool, yielding neat proofs of certain results on division algebras. Here are a few examples: (i) ...
  58. [58]
    [PDF] Four Group-theoretic Proofs of Wedderburn's Little Theorem - OU Math
    Wedderburn proved in 1905 that a finite division ring is always a field. His result has intrigued generations of mathematicians, spurring generalizations and.
  59. [59]
    [PDF] Finite Fields and Function Fields
    Section 1.1 presents some fundamental results on finite fields, such as the existence and uniqueness of finite fields and the fact that the multiplicative group ...
  60. [60]
    [PDF] Algebraic closure
    The algebraic closure of a field K has the same cardinality as K if K is infinite, and is countably infinite if K is finite. Examples. • The fundamental theorem ...
  61. [61]
    [PDF] Fields and Galois Theory - James Milne
    that time not even nineteen years old—defined finite fields of arbitrary prime power order and established their basic properties, e.g. the existence of a ...<|control11|><|separator|>
  62. [62]
    [PDF] Elliptic curves in function field arithmetic
    By a function field, we will mean a function field of one variable over a finite field, i.e., a finite extension of some F(t). Number theory studies number ...
  63. [63]
    [PDF] Curves over finite fields
    We will denote by k(E), the field of fractions of k[E]. This is called the function field of E. Exercise 4. If you haven't seen much algebraic geometry ...
  64. [64]
    [PDF] THE WEIL CONJECTURE. I
    Oct 24, 2021 · In this article, I prove the Weil conjecture on the eigenvalues of Frobenius endomor- phisms. The precise statement is given in (1.6).
  65. [65]
    [PDF] Introduction to Drinfeld modules - MIT Mathematics
    Modularity of elliptic curves over global function fields: If E over Fp(T) has split multi- plicative reduction at ∞, then E is dominated by a Drinfeld modular ...