Fact-checked by Grok 2 weeks ago

Finite group

In , particularly in the field of , a finite group is a group that consists of a finite number of elements under a binary operation satisfying closure, associativity, the existence of an , and inverses for each element. The number of elements in such a group G, denoted |G|, is called the order of the group. Finite groups are essential structures for modeling symmetries in mathematical objects, physical systems, and combinatorial problems, often arising in the study of permutations, rotations, and modular arithmetic. Prominent examples include the cyclic group \mathbb{Z}/p\mathbb{Z} of order p (where p is prime), which is generated by a single element under addition modulo p; the symmetric group S_n, which comprises all permutations of n objects and has order n!; and the alternating group A_n, consisting of even permutations with order n!/2 for n \geq 2. A foundational result, Lagrange's theorem, asserts that if H is a subgroup of a finite group G, then the order of H divides the order of G, implying that the order of any element in G also divides |G|. The theory of finite groups encompasses deep results on their structure and , with simple groups serving as the "atoms" or indecomposable building blocks, unable to be expressed as nontrivial quotients of smaller groups. The (CFSG), one of the most significant theorems in modern , proves that every finite simple group belongs to one of 18 infinite families (such as cyclic groups of prime , alternating groups A_n for n \geq 5, and groups of Lie type) or one of 26 exceptional "sporadic" groups, like the of approximately $8 \times 10^{53}. This , developed over decades by more than 100 mathematicians and spanning over 15,000 pages of proofs, has profound implications for , , and applications in physics, such as particle symmetries.

Fundamentals

Definition

In , a group is a nonempty set G equipped with a \cdot: G \times G \to G that satisfies four axioms: (for all a, b \in G, a \cdot b \in G); associativity (for all a, b, c \in G, (a \cdot b) \cdot c = a \cdot (b \cdot c)); (there exists an element e \in G such that for all a \in G, a \cdot e = e \cdot a = a); and invertibility (for each a \in G, there exists a^{-1} \in G such that a \cdot a^{-1} = a^{-1} \cdot a = e)./02%3A_Groups_I/2.02%3A_Definition_of_a_Group) A finite group is a group whose underlying set G has finite , denoted |G| = n where n is a positive called the of the group. Groups may use multiplicative notation (with \cdot and e) or additive notation (with + and $0), depending on context; for example, the set \mathbb{Z}/n\mathbb{Z} of modulo n under forms a finite group of n./02%3A_Groups_I/2.02%3A_Definition_of_a_Group) The consists of a single element \{e\} satisfying all group axioms, with |G| = 1. Finite groups contrast with infinite groups, where |G| is infinite, though both share the same axiomatic structure.

Order of elements and groups

In a group G with identity element e, the order of an element g \in G, denoted o(g) or |g|, is the smallest positive integer k such that g^k = e, provided such a k exists; otherwise, the order is defined to be . In the context of finite groups, every element has finite order, as the powers of g cannot cycle indefinitely within a set of bounded size. Moreover, in any finite group G of order |G|, the order of every element divides |G|, a property that highlights the structural constraints imposed by finiteness. The cyclic subgroup generated by an element g \in G, denoted \langle g \rangle, consists of all integer powers of g: \langle g \rangle = \{ g^k \mid k \in \mathbb{Z} \}. If g has finite order k, then \langle g \rangle = \{ e, g, g^2, \dots, g^{k-1} \}, and the order of this subgroup equals k, the order of g. This subgroup provides insight into the local structure around g, as its size directly reflects how many distinct powers g produces before returning to the identity. The orders of elements thus reveal much about the overall group structure, with elements of larger orders generating larger cyclic subgroups that embed within G. For example, consider the additive \mathbb{Z}_n of integers n, which has n. The of an m \in \mathbb{Z}_n (with $0 \leq m < n) is n / \gcd(m, n), the smallest positive integer k such that k m \equiv 0 \pmod{n}. Thus, generators like m = 1 have n, while elements sharing factors with n yield smaller orders; for instance, in \mathbb{Z}_{12}, the of 4 is 3 since $3 \cdot 4 = 12 \equiv 0 \pmod{12}. In the symmetric group S_3 of 6, which permutes three , the identity has 1, transpositions like (1\ 2) have 2 (as (1\ 2)^2 = e), and 3-cycles like (1\ 2\ 3) have 3 (as (1\ 2\ 3)^3 = e). These orders—1, 2, and 3—all divide 6, illustrating the general relation in finite groups.

Basic theorems

Lagrange's theorem

Lagrange's theorem asserts that if G is a finite group and H is a subgroup of G, then the order of H, denoted |H|, divides the order of G, denoted |G|. The proof relies on the notion of cosets. For a subgroup H of G, a left coset of H is a set of the form gH = \{ gh \mid h \in H \} where g \in G. A right coset is defined analogously as Hg = \{ hg \mid h \in H \}. Consider the relation \sim on G given by g_1 \sim g_2 if and only if g_1 H = g_2 H (equivalently, g_1^{-1} g_2 \in H). This relation is reflexive, symmetric, and transitive, hence an equivalence relation. The equivalence classes are the distinct left cosets of H, which partition G into disjoint sets. Moreover, any two distinct left cosets are disjoint, and every element of G belongs to exactly one left coset. Each left coset has cardinality |H|, as there is a bijection between H and gH given by left multiplication by g. Let [G : H] denote the number of distinct left cosets of H in G, called the index of H in G. Then G is the disjoint union of these [G : H] cosets, so |G| = [G : H] \cdot |H|. Since [G : H] is a positive integer, it follows that |H| divides |G|. A direct consequence is that the order of any element g \in G divides |G|. Indeed, the cyclic subgroup \langle g \rangle generated by g has order equal to the order of g, and thus divides |G| by the theorem.

Consequences of Lagrange's theorem

Lagrange's theorem implies that the order of every subgroup divides the order of the group, but the converse does not hold in general. For instance, the alternating group A_4 has order 12, yet it contains no subgroup of order 6. A significant structural consequence arises for subgroups of small index. Specifically, any subgroup H of a finite group G with index [G : H] = 2 is normal in G. This follows because the left and right cosets of H coincide, as there are only two cosets: H itself and its complement G \setminus H. Another key implication is Cauchy's theorem, which states that if p is a prime dividing the order of a finite group G, then G contains an element of order p. In the context of number theory, Lagrange's theorem applies to the multiplicative group U(n) of integers modulo n that are coprime to n, which has order \phi(n), where \phi is Euler's totient function. For any g \in U(n), the order of g divides \phi(n), so g^{\phi(n)} \equiv 1 \pmod{n}. This is known as Euler's theorem. When n = p is prime, \phi(p) = p-1, yielding Fermat's little theorem: if p does not divide g, then g^{p-1} \equiv 1 \pmod{p}. This is a special case of Euler's theorem.

Examples

Permutation groups

A permutation group is a subgroup of the symmetric group on a finite set, where the group operation is composition of permutations. The symmetric group S_n, or \mathrm{Sym}(n), consists of all bijections from a set of n elements to itself and has order n!. The alternating group A_n is the subgroup of S_n generated by even permutations, which are those expressible as a product of an even number of transpositions (2-cycles). It has index 2 in S_n, making it a normal subgroup of order n!/2. For n \geq 5, A_n is simple, meaning it has no nontrivial normal subgroups. A transposition is a 2-cycle that swaps two elements while fixing the rest, and every permutation in S_n can be uniquely decomposed (up to ordering) into a product of disjoint cycles, including fixed points as 1-cycles. For example, the symmetric group S_3 has order 6 and is non-abelian, as compositions like (1\ 2)(2\ 3) = (1\ 2\ 3) and (2\ 3)(1\ 2) = (1\ 3\ 2) do not commute. Another example is the dihedral group D_n, which is a permutation group of order $2n realizing the rotational and reflection symmetries of a regular n-gon. Permutation groups also model puzzles like the Rubik's Cube, invented in 1974, where the set of legal moves generates a subgroup of the direct product of symmetric groups on edges and corners, with order approximately $4.3 \times 10^{19}.

Cyclic groups

A cyclic group is a group that can be generated by a single element, known as a generator. For a finite cyclic group G of order n, there exists an element g \in G such that G = \langle g \rangle = \{e, g, g^2, \dots, g^{n-1}\}, where g^n = e is the identity element.) All finite cyclic groups of order n are isomorphic to one another and can be represented additively as the cyclic group \mathbb{Z}/n\mathbb{Z} of integers modulo n under addition. Multiplicatively, they are isomorphic to the group of nth roots of unity in the complex numbers, consisting of solutions to z^n = 1. Every subgroup of a finite cyclic group of order n is itself cyclic, and for each positive divisor d of n, there exists exactly one subgroup of order d, generated by g^{n/d}. A finite group G of order n is cyclic if and only if, for every divisor d of n, G contains exactly \phi(d) elements of order d, where \phi denotes Euler's totient function. Representative examples include the additive group \mathbb{Z}/12\mathbb{Z}, which models clock arithmetic where addition is performed modulo 12 and generated by 1. Another is the group of rotations of a regular n-gon about its center, which is cyclic of order n under composition.

Finite abelian groups

Finite abelian groups extend the structure of cyclic groups by allowing direct products of multiple cyclic components, enabling a complete classification up to isomorphism. Unlike cyclic groups, which are generated by a single element, finite abelian groups can have more complex decompositions while remaining commutative. The key result characterizing these groups is the Fundamental Theorem of Finite Abelian Groups, which states that every finite abelian group G is isomorphic to a direct product of cyclic groups of prime power order:
G \cong \mathbb{Z}/p_1^{k_1}\mathbb{Z} \times \mathbb{Z}/p_1^{k_2}\mathbb{Z} \times \cdots \times \mathbb{Z}/p_m^{k_m}\mathbb{Z} \times \cdots,
where the primes p_i may repeat and the exponents k_j are positive integers. This theorem provides a canonical form that uniquely determines the group's isomorphism class based on its order.
The primary decomposition aspect of the theorem decomposes G into its Sylow p-subgroups for each prime p dividing |G|, yielding
G \cong \bigoplus_p G_p,
where each G_p is the p-primary component, a finite abelian p-group isomorphic to a direct sum of cyclic groups of orders powers of p: G_p \cong \mathbb{Z}/p^{a_1}\mathbb{Z} \oplus \mathbb{Z}/p^{a_2}\mathbb{Z} \oplus \cdots \oplus \mathbb{Z}/p^{a_r}\mathbb{Z} with a_1 \geq a_2 \geq \cdots \geq a_r \geq 1. These components are independent, as the order of G factors into distinct prime powers. The elementary divisors form refers to the collection of these prime power orders p^{a_i}, which fully specify the group up to isomorphism when sorted appropriately.
In contrast, the invariant factors decomposition expresses G as a direct product of cyclic groups \mathbb{Z}/m_1\mathbb{Z} \times \mathbb{Z}/m_2\mathbb{Z} \times \cdots \times \mathbb{Z}/m_s\mathbb{Z}, where m_1 divides m_2, m_2 divides m_3, and so on, up to m_s, and the product of the m_i equals |G|. This form is unique and often more compact for computation, though deriving it from the elementary divisors involves combining prime powers across different primes while preserving the divisibility condition. Both decompositions are equivalent representations of the same theorem, with transformations between them possible via prime factorization. A representative example is the Klein four-group V_4, which has order 4 and is isomorphic to \mathbb{Z}/2\mathbb{Z} \times \mathbb{Z}/2\mathbb{Z}. This group is non-cyclic, as no single element generates it—all non-identity elements have order 2—and its primary decomposition consists of two copies of the . In invariant factors form, it remains \mathbb{Z}/2\mathbb{Z} \times \mathbb{Z}/2\mathbb{Z}, since 2 divides 2. Another illustration is the abelian group of order 8 given by \mathbb{Z}/2\mathbb{Z} \times \mathbb{Z}/4\mathbb{Z}, with elementary divisors 2 and 4, contrasting with the \mathbb{Z}/8\mathbb{Z}. Homomorphisms between finite abelian groups G and H are determined by the primary decompositions: \Hom(G, H) \cong \prod_p \Hom(G_p, H_p), where each component homomorphism respects the cyclic sum structure. Endomorphisms of G, i.e., \End(G), form a ring whose structure mirrors the direct sum of matrix rings over the integers modulo the relevant orders, facilitating computations of the group's module-like properties.

p-groups

A finite p-group is a finite group G whose order |G| is a power of a prime p, that is, |G| = pk for some integer k ≥ 1. Every nontrivial finite p-group has a nontrivial Z(G), meaning |Z(G)| ≥ p. Moreover, every maximal of a finite p-group is and has p. A key structural property is that every finite p-group of pk possesses a of pm for each m with 1 ≤ mk; in fact, the number of such subgroups depends on the group's structure, but their existence follows from iteratively applying the nontrivial center property to build a chain of normal subgroups. Prominent examples include the quaternion group Q8 of order 8 (p=2), which is non-abelian and has presentation <a,b | a4=1, a2=b2, b-1a(b)=a-1>; its center is <a2> of order 2, and it has three subgroups of order 4, all normal and cyclic. Another example is the elementary abelian p-group (Z/pZ)k, which is abelian, isomorphic to the vector space (Fp)k under addition, where every non-identity element has order p and all subgroups are normal. The Frattini subgroup Φ(G) of a finite p-group G is the intersection of all its maximal subgroups, which coincides with the subgroup generated by all commutators [g,h] and all pk-th powers gpk for gG; notably, G/Φ(G) is elementary abelian of order pd, where d is the minimal number of generators of G. Finite p-groups admit a chief series, a maximal normal series 1 = N0N1 ⊴ ⋯ ⊴ Nk = G where each factor N*i+1/Ni is a minimal normal subgroup of G/Ni and has order p, reflecting the p-group's nilpotent structure with chief factors that are elementary abelian of rank 1. The Burnside basis theorem states that if G is a finite p-group, then the minimal number of generators d(G) equals the dimension of the elementary abelian group G/Φ(G) over Fp, and any generating set of G maps onto a basis of G/Φ(G) if and only if it generates G.

Groups of Lie type

Groups of Lie type are finite groups constructed as the fixed points under certain endomorphisms of algebraic groups defined over finite fields, serving as discrete analogues of classical Lie groups. These groups emerge from the rational points of a reductive algebraic group G defined over a finite field \mathbb{F}_q, where q is a power of a prime, typically taking the form G(\mathbb{F}_q) or subgroups thereof, such as the special linear group \mathrm{SL}(n, q) or its projective version \mathrm{PSL}(n, q). The foundational construction relies on the structure of semisimple Lie algebras over the complex numbers, transported to characteristic p via a Chevalley basis, which allows the generation of these finite groups uniformly across different types. The untwisted groups of Lie type, known as Chevalley groups, are classified by Dynkin diagrams corresponding to simple algebras and include families such as A_n(q) = \mathrm{PSL}(n+1, q), the projective special linear groups; B_n(q), the odd-dimensional orthogonal groups over \mathbb{F}_q; C_n(q), the groups \mathrm{Sp}(2n, q); D_n(q), the even-dimensional orthogonal groups; and the exceptional types E_6(q), E_7(q), E_8(q), F_4(q), and G_2(q). Twisted variants, introduced to capture additional simple groups, arise by applying a (field automorphism combined with a ) to the Chevalley group; prominent examples are the groups ^2B_2(q) for q = 2^{2m+1}, the Ree groups of type ^2G_2(q) for q = 3^{2m+1}, and ^2F_4(q) for q = 2^{2m+1}. These twisted groups fill out the complete of non-abelian finite simple groups of Lie type, excluding the alternating and sporadic families. The orders of these groups follow explicit formulas derived from the and structure. For instance, the order of \mathrm{PSL}(2, q) is \frac{q(q-1)(q+1)}{d}, where d = \gcd(2, q-1), reflecting the quotient of \mathrm{SL}(2, q) by its ; this yields q(q^2 - 1)/2 when q is odd and q(q^2 - 1) when q is a power of 2. Similar expressions hold for higher-rank groups, scaling with q raised to the of the variety, modulated by factors from the and order. These formulas underscore the groups' non-abelian nature and their role as rich examples beyond cyclic or abelian structures. In the , groups of Lie type constitute the largest infinite families, comprising 16 series (including twists) that account for the majority of all known non-abelian groups, with the remainder being alternating groups, cyclic groups of prime order, and sporadics. This prominence stems from their systematic , which unifies diverse linear and orthogonal groups under . The theory originated with Claude Chevalley's 1955 of the untwisted groups via integral forms of Lie algebras, later extended by Robert Steinberg in the late through twisting mechanisms to encompass the full spectrum of Lie-type simples.

Sylow theory

Sylow theorems

A Sylow p-subgroup of a finite group G is a maximal p- of G, meaning a P \leq G whose is p^k, where p^k is the highest power of the prime p dividing |G|. Such subgroups play a central role in the structure theory of finite groups, generalizing the concept of p-groups to arbitrary finite groups. The first Sylow theorem guarantees the existence of such subgroups. Sylow's first theorem states that for a finite group G and a prime p dividing |G|, there exists at least one Sylow p- of G. The proof proceeds by on the order of G, building larger p-subgroups step by step. If |G| is a power of p, then G itself is the Sylow p-subgroup. Otherwise, start with a nontrivial p-subgroup H and consider its action on the left cosets of itself by left multiplication. The fixed-point congruence implies that the normalizer N_G(H) properly contains H, allowing construction of a larger p-subgroup by induction or Cauchy's theorem. Continuing this process yields a maximal p-subgroup of order p^k. The second Sylow theorem addresses the conjugacy of these subgroups. Sylow's second theorem asserts that any two Sylow p-subgroups of G are conjugate in G, and moreover, the number n_p of distinct Sylow p-subgroups satisfies n_p \equiv 1 \pmod{p} and divides |G|/p^k. The conjugacy follows from the transitivity of the conjugation action of G on the set of Sylow p-subgroups. For a fixed Sylow p-subgroup P, the under this action is N_G(P), so n_p = [G : N_G(P)], which divides |G|/p^k by (since |N_G(P)| is divisible by p^k). The congruence n_p \equiv 1 \pmod{p} arises from the action of P on the left cosets of N_G(P) by left multiplication: since P \leq N_G(P), this action fixes exactly one coset (the trivial one), and fixed-point theorems imply the number of cosets (i.e., n_p) is congruent to 1 modulo p. Finally, Sylow's third theorem provides a criterion for normality: a Sylow p-subgroup P of G is in G if and only if it is the unique Sylow p-subgroup of G. The forward direction follows immediately from the second theorem, as conjugates of a subgroup are itself. Conversely, if P is unique, then it is fixed by all conjugations, hence . These theorems, originally proved by Peter Ludvig Sylow in 1872, form the foundation for much of modern finite group theory.

Applications of Sylow theorems

The Sylow theorems provide powerful tools for classifying finite groups of prime-power product order, particularly when the order is pq with distinct primes p < q. In such a group G, the number of Sylow q-subgroups n_q divides p and satisfies n_q \equiv 1 \pmod{q}. Since p < q, the only possibility is n_q = 1, so the Sylow q- is unique and hence normal in G. The number of Sylow p-subgroups n_p divides q and satisfies n_p \equiv 1 \pmod{p}, so n_p = 1 or q. The case n_p = q occurs if and only if q \equiv 1 \pmod{p}, or equivalently, p divides q-1. If n_p = 1, then both Sylow subgroups are normal, and G is cyclic of order pq. If n_p = q > 1, then G is a non-trivial of its normal Sylow q-subgroup by the Sylow p-subgroup. A key consequence of the concerns solvability: if every Sylow subgroup of a finite group G is , then G is the of its Sylow subgroups. This follows because the Sylow subgroups pairwise normalize each other, and their elements commute across distinct primes, yielding a direct decomposition into cyclic p-groups for each prime p dividing |G|. Such groups are necessarily solvable, as the direct product of solvable groups (here, cyclic p-groups) is solvable. The facilitate this by confirming the uniqueness and normality of these subgroups via n_p = 1 for all primes p. The classification of groups of order 12 illustrates practical applications of Sylow counts to determine group structure. For |G| = 12 = 2^2 \cdot 3, the possible values are n_3 = 1 or $4 (dividing 4 and \equiv 1 \pmod{3}) and n_2 = 1 or $3 (dividing 3 and \equiv 1 \pmod{2}). If n_3 = 1, the normal Sylow 3-subgroup Q \cong \mathbb{Z}_3 admits either a (yielding \mathbb{Z}_{12} or \mathbb{Z}_6 \times \mathbb{Z}_2) or a by a Sylow 2-subgroup (yielding the D_{12} of order 12 or the of order 12). If n_3 = 4, then n_2 = 1 or $3; the case n_2 = 3 is impossible as it would imply more than 12 elements, so n_2 = 1 and G \cong A_4, the on 4 letters. These cases exhaust the five isomorphism classes of groups of order 12. The Sylow theorems also transfer to the study of composition factors by helping identify subgroups and in a . For instance, a Sylow p- yields a that is a p-complement, allowing recursive decomposition; non-trivial Sylow counts can signal non-solvability or specific simple factors, as in the where Sylow subgroups constrain possible orders. Burnside's normal p-complement theorem provides a for the existence of a normal Hall p'-. Specifically, if G is finite and P is a Sylow p- with P \leq Z(N_G(P)), then G has a normal p-complement N (a normal Hall of order |G|/|P|) such that G = N \rtimes P. This condition leverages the by ensuring the normalizer controls fusion and centralization within P, often verified via n_p \equiv 1 \pmod{p} and additional divisibility. The theorem implies solvability in cases where such complements exist for the smallest prime p dividing |G|.

Group actions

Cayley's theorem

states that every finite group G is isomorphic to a of the \Sym(G) consisting of all bijections from G to itself. This isomorphism arises from the left regular of G on itself, defined by g \cdot x = gx for all g, x \in G. To establish this, consider the map \phi: G \to \Sym(G) given by \phi(g)(x) = gx. First, \phi(g) is a for each g \in G: it is injective because if g x_1 = g x_2, then left multiplication by g^{-1} yields x_1 = x_2; it is surjective because for any x \in G, there exists x' = g^{-1} x such that \phi(g)(x') = x. Next, \phi is a : \phi(g_1 g_2)(x) = (g_1 g_2) x = g_1 (g_2 x) = \phi(g_1)(\phi(g_2)(x)), so \phi(g_1 g_2) = \phi(g_1) \circ \phi(g_2). Finally, \phi is injective (hence faithful), as \phi(g_1) = \phi(g_2) implies \phi(g_1)(e) = \phi(g_2)(e), so g_1 = g_2 where e is the . Thus, \phi embeds G as a of \Sym(G), which has degree |G|. The implies that every finite group can be realized concretely as a acting regularly on a set of size equal to its , providing a bridge from abstract algebraic structures to explicit symmetries of finite sets. This realization underscores the connection between and the study of symmetries, as permutation groups model transformations preserving set structure. For example, consider the S_3 of 6, generated by a 3- f = (1\ 2\ 3) and a g = (1\ 2). The left regular action embeds S_3 into \Sym(S_3), where elements act by left multiplication on the group's own elements (listed as e, f, f^2, g, fg, f^2 g). For instance, \phi(f) permutes these as e \mapsto f, f \mapsto f^2, f^2 \mapsto e, g \mapsto fg, fg \mapsto f^2 g, f^2 g \mapsto g, corresponding to the (e\ f\ f^2)(g\ fg\ f^2 g). This permutation faithfully captures S_3's within the larger symmetric group of 6.

Burnside's lemma

In group theory, a finite group G acts on a finite set X if there is a map G \times X \to X, denoted (g, x) \mapsto g \cdot x, such that the fixes every point and the action is compatible with the group operation: e \cdot x = x and (gh) \cdot x = g \cdot (h \cdot x) for all g, h \in G and x \in X. The of an element x \in X is the set \{ g \cdot x \mid g \in G \}, which partitions X into equivalence classes under the relation x \sim y if y = g \cdot x for some g \in G. The of x is the \operatorname{Stab}_G(x) = \{ g \in G \mid g \cdot x = x \}. Burnside's lemma provides a method to count the number of orbits in such an . For a finite group G acting on a X, the number of orbits is given by \frac{1}{|G|} \sum_{g \in G} |\operatorname{Fix}(g)|, where \operatorname{Fix}(g) = \{ x \in X \mid g \cdot x = x \} is the set of fixed points of g. This formula, originally attributed to Frobenius but popularized by Burnside, arises from averaging the number of fixed points over all group elements. To sketch the proof, consider the sum \sum_{g \in G} |\operatorname{Fix}(g)|, which equals \sum_{x \in X} |\operatorname{Stab}_G(x)| by double counting the pairs (g, x) with g \cdot x = x. For each orbit O, the stabilizers of its elements are equal, and by the orbit-stabilizer theorem, |O| = |G| / |\operatorname{Stab}_G(x)| for x \in O, so \sum_{x \in O} |\operatorname{Stab}_G(x)| = |O| \cdot |\operatorname{Stab}_G(x)| = |G|. Summing over all orbits thus yields \sum_{x \in X} |\operatorname{Stab}_G(x)| = |G| \cdot k, where k is the number of orbits, proving the lemma. Burnside's lemma has key applications in finite group theory, such as counting necklaces under the action of the of rotations, where elements with cycle structures matching the necklace's symmetries contribute to fixed colorings. It also enumerates conjugacy classes in G by applying the lemma to the conjugation action on G itself, yielding the number of classes as \frac{1}{|G|} \sum_{g \in G} |\operatorname{C}_G(g)|, where \operatorname{C}_G(g) is the centralizer of g. Additionally, it counts conjugacy classes of subgroups, providing the number of subgroups up to under conjugation. A representative example is counting the number of distinct colorings of an n-element set with k colors up to permutation by the S_n. The set X consists of all functions from \{1, \dots, n\} to \{1, \dots, k\}, with S_n acting by (g \cdot f)(i) = f(g^{-1} i). A g fixes a coloring f if f is constant on the cycles of g, so |\operatorname{Fix}(g)| = k^{c(g)} where c(g) is the number of cycles in g. The number of orbits is thus \frac{1}{n!} \sum_{g \in S_n} k^{c(g)}.

Structure theorems

Direct and semidirect products

The direct product of two finite groups G and H, denoted G \times H, consists of ordered pairs (g, h) with g \in G and h \in H, equipped with the componentwise operation (g_1, h_1)(g_2, h_2) = (g_1 g_2, h_1 h_2). This construction yields a group of |G| \cdot |H|, and the projections onto each factor are surjective homomorphisms with kernels isomorphic to the other factor. If both G and H are abelian, then G \times H is abelian, since for any (g_1, h_1), (g_2, h_2) \in G \times H, the [(g_1, h_1), (g_2, h_2)] = ([g_1, g_2], [h_1, h_2]) = (e_G, e_H). An internal direct product characterizes when a finite group G decomposes as such a product of its subgroups: G = N \times K if and only if N and K are normal subgroups of G, N \cap K = \{e\}, and N K = G. In this case, every element of G uniquely writes as n k with n \in N and k \in K, and the multiplication follows the rule. A representative example is the , which is the direct product \mathbb{Z}/2\mathbb{Z} \times \mathbb{Z}/2\mathbb{Z}, consisting of elements of order dividing 2 under componentwise addition modulo 2. The provides a more general construction for building , incorporating a nontrivial action. Given N and H and a \phi: H \to \Aut(N), the external N \rtimes_\phi H has underlying set N \times H with operation (n_1, h_1)(n_2, h_2) = (n_1 \cdot \phi(h_1)(n_2), h_1 h_2). This forms a group where N (identified with N \times \{e_H\}) is and H (identified with \{e_N\} \times H) is a , with N \cap H = \{e\} and N H = N \rtimes_\phi H. Internally, G = N \rtimes H if N is in G, H is a , N \cap H = \{e\}, and N H = G, with conjugation in G inducing the action \phi. A classic example is the S_3, which is the \mathbb{Z}/3\mathbb{Z} \rtimes \mathbb{Z}/2\mathbb{Z}, where \mathbb{Z}/2\mathbb{Z} acts on \mathbb{Z}/3\mathbb{Z} by inversion (the nontrivial sending $1 \mapsto 2 \pmod{3}). Here, the order-3 rotation subgroup is normal, and the order-2 reflection complements it. The arises as a special case of the when \phi is the trivial , yielding no twisting by automorphisms.

Solvable and nilpotent groups

A is a finite group G that possesses a subnormal series \{e\} = G_0 \trianglelefteq G_1 \trianglelefteq \cdots \trianglelefteq G_k = G such that each factor group G_{i+1}/G_i is abelian. Equivalently, the derived series of G, defined by G^{(0)} = G and G^{(i+1)} = [G^{(i)}, G^{(i)}] (the ), terminates at the trivial after finitely many steps, i.e., G^{(n)} = \{e\} for some n. This property captures groups that can be "built up" from abelian groups through extensions, reflecting a hierarchical structure amenable to inductive analysis. All abelian groups are solvable, as their derived subgroup is trivial. For instance, the symmetric group S_3 of order 6 is solvable, with derived series S_3 \triangleright A_3 \triangleright \{e\}, where A_3 is cyclic of order 3. In contrast, the alternating group A_5 of order 60 is not solvable, as its derived subgroup equals itself, preventing the series from reaching the trivial group. A finite group is if its lower central series, defined by \gamma_1(G) = G and \gamma_{i+1}(G) = [G, \gamma_i(G)], terminates at the trivial , i.e., \gamma_m(G) = \{e\} for some m. For finite groups, this is equivalent to the group being the of its Sylow subgroups, each of which is . groups form a subclass of solvable groups, as the lower central series refines to a subnormal series with abelian factors. Every finite p-group is nilpotent (and hence solvable), since the center of a nontrivial finite p-group is nontrivial, allowing the upper central series to ascend to the whole group in finitely many steps. Abelian groups are nilpotent of class 1, with trivial lower central series beyond the first term. The group S_3 is solvable but not nilpotent, as its lower central series stabilizes at A_3 \neq \{e\}. Burnside's normal p-complement theorem provides a criterion for solvability: if P is a Sylow p- of a finite group G such that P lies in the center of its normalizer N_G(P), then G has a normal p-complement (a normal Hall whose order is coprime to p and intersects P trivially). Iteratively applying this theorem to the factors can establish solvability, as the existence of such complements reduces the problem to smaller solvable pieces.

Composition series and Jordan–Hölder theorem

A composition series of a finite group G is a finite chain of subgroups $1 = G_0 \trianglelefteq G_1 \trianglelefteq \cdots \trianglelefteq G_n = G such that each quotient G_{i+1}/G_i is a simple group for $0 \leq i < n; these quotients are called the composition factors of the series. Every finite group possesses at least one composition series, which can be constructed by iteratively selecting maximal normal subgroups until reaching the trivial subgroup. The Jordan–Hölder theorem asserts that any two of a finite group G have the same length n and the same composition factors up to and . This uniqueness implies that the of composition factors is an invariant of the group, providing a into building blocks. The proof relies on the Schreier refinement theorem, which states that any two subnormal series of a group admit refinements that are equivalent, meaning their factor groups are isomorphic up to permutation and repetition. To apply this to , one refines both series using the to ensure maximal subnormal steps with factors, then removes isomorphic repetitions to match the factors pairwise; the process uses induction on the group order to handle the base case of groups. A related concept is the series, a maximal chain of s $1 = N_0 \trianglelefteq N_1 \trianglelefteq \cdots \trianglelefteq N_r = G where each N_{i+1}/N_i is a minimal of G/N_i, known as chief factors; unlike composition factors, chief factors need not be but are characteristically , often direct products of isomorphic groups. The Jordan–Hölder theorem extends analogously to chief series, ensuring their factors are unique up to and . For example, the S_4 has chief series \{e\} \trianglelefteq V_4 \trianglelefteq A_4 \trianglelefteq S_4, where V_4 is the , with chief factors \mathbb{Z}_2 \times \mathbb{Z}_2, \mathbb{Z}_3, and \mathbb{Z}_2. A corresponding refines the nonsimple chief factor: \{e\} \trianglelefteq \langle (1\,2)(3\,4) \rangle \trianglelefteq V_4 \trianglelefteq A_4 \trianglelefteq S_4, yielding factors \mathbb{Z}_2, \mathbb{Z}_2, \mathbb{Z}_3, \mathbb{Z}_2. In solvable groups, all composition factors are abelian, specifically cyclic of prime order.

Simple groups

Definition and basic properties

In group theory, a finite is defined as a nontrivial finite group that possesses no subgroups other than the trivial and the group itself. This definition, originally proposed by , captures groups that cannot be decomposed nontrivially via subgroups, making them the "atoms" of finite group structure. For abelian groups, simplicity implies that the group is cyclic of prime order. Specifically, if G is an abelian , then |G| = p for some prime p, and G \cong \mathbb{Z}/p\mathbb{Z}. This follows from the fact that any proper nontrivial of an abelian group is , so simplicity requires no such subgroups, which occurs precisely when the order is prime. Non-abelian s, by contrast, are infinite in number and include examples like the A_n for n \geq 5, which is simple because any must contain 3-cycles and thus generate the entire group. A key property of simple groups is their role in composition series: every finite group has a composition series where the successive quotients (composition factors) are groups, and by the Jordan–Hölder theorem, these factors are unique up to and ordering. For a G, the only maximal normal subgroup is the trivial subgroup \{e\} (since G itself is normal in G), emphasizing its indecomposability. Although all known non-abelian finite simple groups have even order—with A_5 (order 60) as the smallest example—early conjectures sometimes questioned this, but counterexamples like A_5 confirm their existence. Regarding automorphisms, the outer automorphism group \mathrm{Out}(G) = \mathrm{Aut}(G)/\mathrm{Inn}(G) measures symmetries beyond inner ones induced by conjugation; for most finite simple groups, \mathrm{Out}(G) is small, often of order 1 or 2. The M(G) = H_2(G, \mathbb{Z}) of a finite G is the kernel of the universal central extension and is typically small or trivial for non-abelian cases; for instance, it is trivial for alternating groups A_n (n \geq 5) and cyclic groups \mathbb{Z}_p. This multiplier encodes stem extensions and has been computed for all known finite simple groups, aiding their .

Feit–Thompson theorem

The states that every finite group of odd order is solvable. This result, also known as the odd order theorem, was established by Walter Feit and John G. Thompson in their seminal 1963 paper "Solvability of groups of odd order," published in the Pacific Journal of Mathematics. The proof spans 255 pages and represents one of the most intricate arguments in finite group theory at the time. The proof strategy is divided into local and global components, relying heavily on advanced techniques from . The local analysis employs of finite groups, particularly the use of transfers—maps that relate characters of a group to those of its —to investigate the structure of Sylow subgroups and detect nilpotency in certain formations. Formation theory, a framework for constructing groups via subnormal series with specified factor groups, is then applied in the global phase to show that a minimal must possess a solvable subgroup, leading to a contradiction. Subsequent simplifications, such as those by in the 1970s, have reduced the length while preserving the core ideas of character-theoretic transfers and formations. A key corollary of the theorem is that all non-abelian simple finite groups have even order, since a non-abelian simple group of odd order would contradict solvability while violating simplicity. This implication was pivotal in the classification of finite simple groups, as it eliminated the need to consider odd-order candidates beyond cyclic groups of prime order, thereby focusing efforts on even-order cases and serving as the foundational step in the decades-long project completed in the 1980s and 2000s.

Classification of finite simple groups

The Classification of Finite Simple Groups (CFSG) is one of the most significant achievements in modern , providing a complete enumeration of all finite s up to . This theorem asserts that every finite falls into one of four categories: cyclic groups of prime order, alternating groups A_n for n \geq 5, groups of Lie type defined over finite fields, or one of 26 exceptional sporadic groups. The classification encompasses 18 infinite families in total (including the cyclic and alternating ones within the broader count) and these 26 sporadics, with no others existing. The abelian simple groups are exactly the cyclic groups \mathbb{Z}_p where p is prime. The non-abelian infinite families consist of the alternating groups A_n (n \geq 5), which are the even permutations on n letters, and the 16 families of groups of type. These -type groups arise as finite analogues of groups and include Chevalley groups such as the projective linear groups \mathrm{PSL}(n, q), groups \mathrm{PSp}(2m, q), and exceptional types like E_8(q), all defined over the \mathbb{F}_q where q is a ; twisted , such as the unitary groups \mathrm{PSU}(n, q), Suzuki groups \mathrm{Sz}(q) for q = 2^{2m+1}, and Ree groups {}^2G_2(q) or {}^2F_4(q). The 26 sporadic simple groups are finite exceptions that do not belong to any infinite family and were discovered individually through various constructions. Notable examples include the M_{11}, M_{12}, M_{22}, M_{23}, and M_{24}, which are highly symmetric permutation groups related to Steiner systems; the Janko groups J_1, J_2, J_3, and J_4; the Conway groups \mathrm{Co}_1, \mathrm{Co}_2, and \mathrm{Co}_3, linked to symmetries; and the \mathbb{M}, the largest sporadic with order $8,089,174,247,945,128,785,886,459,904,961,710,757,005,754,368,000,000,000 \approx 8 \times 10^{53}. Twenty of these sporadics are subquotients of the Monster (the "Happy Family"), while the remaining six are "pariahs" with no such connections. The proof of the CFSG involved over 100 mathematicians and spanned more than 50 years, culminating in over 10,000 pages across hundreds of papers; it was initially announced as complete in 1983 by Daniel Gorenstein but required revisions, with the final gaps closed in 2004 by Michael Aschbacher and Stephen D. Smith. Ongoing projects, including a second-generation proof by Gorenstein, Lyons, and Solomon, aim to streamline and verify the result further. A key implication is that, by the Jordan–Hölder theorem, every finite group admits a whose factors are these simple groups, allowing all finite groups to be understood as "built" from them via group extensions, direct products, and semidirect products.

Enumeration

Number of groups of order n

The number g(n) of groups of order n up to , also denoted f(n) in some literature, counts the distinct classes of finite groups with exactly n elements. This function is multiplicative in a certain sense but highly irregular, with g(n) = 1 for all n \leq 3 (the for n=1, and the cyclic groups \mathbb{Z}/2\mathbb{Z} and \mathbb{Z}/3\mathbb{Z} for n=2,3), and it grows rapidly thereafter, particularly when n is highly composite, reflecting the increasing complexity of group structures as more prime factors are introduced. For instance, the proliferation arises from combinations of Sylow subgroups and extensions, leading to an explosion in possibilities for orders with many small prime factors. For specific cases, explicit formulas exist. When n = p^k is a prime power, the enumeration of p-groups of order p^k is a central problem, with the asymptotic growth given by Higman's formula: g(p^k) = p^{\frac{2}{27} k^3 + O(k^{5/2})}. This reflects the polynomial-in-p nature of the count in the exponent of k, driven by the variety of nilpotent structures and relations in p-groups. For the subclass of abelian groups of order n, the fundamental theorem of finite abelian groups provides a complete classification up to isomorphism via invariant factors or elementary divisors, yielding g_{\text{abelian}}(n) = \prod_p p(k_p), where the product is over primes p dividing n, k_p = v_p(n) is the p-adic valuation, and p(m) denotes the partition function counting integer partitions of m. In general, no closed-form formula for g(n) exists, but computational methods enable determination for moderate n. Systems like the (Groups, Algorithms, Programming) package include the SmallGroups library, which catalogs all classes of groups up to 2000 (excluding orders 1024 and 1536 due to computational intensity), facilitating enumeration, , and via algorithms for Sylow subgroups and presentations. Online databases built on such libraries, including those integrated with , provide accessible lookups and verify for research. Asymptotically, bounds on g(n) capture the explosive growth without exact formulas. Pyber established an upper bound g(n) \leq n^{\left( \frac{2}{27} + o(1) \right) \mu(n)^2}, where \mu(n) = \max_p v_p(n) is the largest exponent in the prime of n, implying \log g(n) < \left( \frac{2}{27} + o(1) \right) \mu(n)^2 \log n. These polynomial-exponential bounds highlight that g(n) is subexponential in n, with the dominant contribution often from p-groups for small primes like p=2, aligning with lower bounds from the Higman-Sims asymptotic that suggest \log g(n) grows on the order of (\log n)^3 for prime-power n.

Groups of small order

The only group of order 1 is the . For a p, there is exactly one group of order p up to : the \mathbb{Z}_p. Groups of order p^2, where p is prime, are all abelian; there are exactly two up to isomorphism: the \mathbb{Z}_{p^2} and the \mathbb{Z}_p \times \mathbb{Z}_p. For order pq with distinct primes p < q, the classification depends on the divisibility condition p \mid (q-1). If p does not divide q-1, the only group is the \mathbb{Z}_{pq}. If p divides q-1, there are exactly two groups: the \mathbb{Z}_{pq} and a non-abelian \mathbb{Z}_q \rtimes \mathbb{Z}_p. This uses to show the Sylow q-subgroup is and the action of \mathbb{Z}_p on it is determined by homomorphisms to \mathrm{Aut}(\mathbb{Z}_q) \cong \mathbb{Z}_{q-1}^\times. For example, for order 6 = 2 × 3 (where 2 divides 3-1), there are two groups: \mathbb{Z}_6 and S_3. Although order 12 = 2^2 × 3 is not of the form pq, there are five groups up to : the abelian ones \mathbb{Z}_{12} and \mathbb{Z}_6 \times \mathbb{Z}_2; and the non-abelian ones A_4, the D_{12} of order 12 (symmetries of regular ), and the \mathrm{Dic}_3 (also known as the binary dihedral group of order 12). These are classified using Sylow subgroups and semidirect products, with the non-abelian examples arising from actions of Sylow 3-subgroups on Sylow 2-subgroups or vice versa. The numbers of groups of small order are tabulated below for n \leq 60, including the count of non-abelian groups. These enumerations stem from systematic computational constructions verifying all possibilities up to .
Order nTotal groupsNon-abelian groups
110
210
310
420
510
621
710
852
920
1021
1110
1253
1310
1421
1510
16149
1710
1853
1910
2053
2121
2221
2310
241512
2520
2621
2752
2842
2910
3043
3110
325144
3310
3421
3510
361410
3710
3821
3921
401411
4110
4265
4310
4442
4520
4621
4710
485247
4920
5053
5110
5296
5310
541512
5521
561310
5721
5821
5910
601311
All finite groups of order less than 60 have been explicitly classified up to using these enumerative methods. A notable pattern is the predominance of abelian groups for prime-power orders, with non-abelian examples emerging first at order 6 and increasing rapidly for highly composite orders like , , and 60. In contrast, the classification becomes more complex at order ($2^6), where there are 267 groups, mostly non-abelian p-groups.

History

Early developments

The study of finite groups originated in the through investigations into the symmetries of geometric objects and the structure of equations. Leonhard Euler, in his work on polyhedra during the 1750s and 1760s, examined the rotational symmetries of regular polyhedra such as the platonic solids, implicitly dealing with finite sets of transformations that preserved their forms; these provided early concrete examples of what would later be recognized as finite groups. In the 1770s, advanced this area by analyzing permutations of the roots of polynomial equations in his memoir Réflexions sur la résolution algébrique des équations (1770–1771), where he explored how rearrangements of roots relate to solving equations by radicals, effectively studying the action of the on the roots without explicitly defining the group structure. This approach highlighted the finite nature of permutation sets and their role in algebraic solvability. Paolo Ruffini built on these ideas in 1799 with his Teoria generale delle equazioni, in which he proved that general polynomial equations of degree five or higher cannot be solved by radicals, using arguments involving the order and properties of permutations of roots; this result, known as Ruffini's theorem, was the first major demonstration of the limitations of radical solutions and anticipated key aspects of group-theoretic solvability. Augustin-Louis Cauchy formalized early group concepts in 1812 through his memoir on symmetric functions and permutations, submitted to the French Academy and published in 1815, where he introduced the idea of a "group of substitutions" as a of permutations, along with results on their orders and cycles that prefigured abstract . , in the early 1830s, revolutionized the field by associating to each its —a finite group of permutations of the roots that encodes the symmetries of the equation's —showing in memoirs submitted to the Academy in 1830 and 1831 that solvability by radicals corresponds precisely to the group being solvable. Galois's insights, though not fully appreciated until their posthumous publication in 1846, provided the abstract framework linking finite groups to the resolvability of equations.

19th-century advances

The full development of , which laid foundational insights into the structure of finite permutation groups and their relation to solvability of equations, occurred posthumously through the publication of Évariste Galois's manuscripts in 1846 by in the Journal de Mathématiques Pures et Appliquées. This edition compiled Galois's earlier unpublished works, including analyses of group actions on roots, establishing key correspondences between subgroups and field extensions that influenced subsequent finite group studies. In 1854, introduced the first abstract definition of a group, conceptualizing it as a set of symbols satisfying certain associative laws under a , independent of specific realizations like permutations or matrices. This shift from concrete representations to abstract structures enabled broader applications in finite group theory. Cayley also proved that every finite group is isomorphic to a of the on its elements, a result now known as . Camille Jordan advanced the understanding of finite group decompositions in his 1870 treatise Traité des substitutions et des équations algébriques, where he introduced the concept of as a chain of normal subgroups with simple factor groups. Jordan's work demonstrated that such series provide invariant structural information about solvable groups, building on Galois's ideas to analyze permutation groups systematically. In 1884, integrated finite groups into geometric contexts, notably applying the icosahedral rotation group to resolve the general quintic equation via modular functions and symmetries of the . Klein's approach highlighted how finite groups govern transformations in non-Euclidean geometries, as explored in his of 1872, which classified geometries by their underlying symmetry groups. Peter Ludvig Sylow's 1872 theorems provided crucial tools for dissecting finite groups by prime powers, stating that for a prime p dividing the group order, Sylow p-subgroups exist, are conjugate, and their number satisfies specific congruence conditions. These results facilitated the study of structures and influenced early attempts for groups of small orders. Nineteenth-century efforts also included initial classifications of dyadic groups—finite 2-groups—and broader enumerations of groups up to certain orders, often leveraging to identify classes, as seen in works by mathematicians like and later compilers of tables for orders through 100. These endeavors marked the transition toward systematic catalogs, though complete listings remained elusive until later refinements.

20th-century milestones

In the early 1900s, William Burnside made significant advances in the study of finite groups, notably posing the in 1902, which questions whether a in which every element has bounded finite order must itself be finite. This problem, arising from observations on periodic groups, spurred extensive research into torsion groups and their finiteness properties, influencing later work on solvable groups. Burnside also proved in 1904 that any finite group of order p^a q^b, where p and q are distinct primes and a, b are non-negative integers, is solvable, a result that relies on and marked a key step toward understanding solvability for groups with few prime factors. This theorem provided early evidence that non-solvable finite groups require more complex structures in their orders. Around the same period, advanced the , developing foundational tools in the early 1900s that linked group structure to linear algebra over the complex numbers. His work, including the introduction of in 1901, established that endomorphisms of irreducible representations are scalars, enabling the decomposition of representations into irreducibles and facilitating applications to group characters and solvability criteria. Schur's contributions, such as proofs of the integrality of characters and orthogonality relations, became essential for analyzing finite group symmetries and were instrumental in later classification efforts. In the 1950s, introduced the Chevalley groups, providing a uniform construction of finite simple groups of Lie type over finite fields, which form one of the infinite families in the . These groups, defined via root systems and Chevalley bases for semisimple Lie algebras, include analogues of classical groups like PSL(n,q) and exceptional types, and their development in works such as Chevalley's 1955 seminar notes marked a shift toward in finite group theory. This framework clarified the structure of Lie-type groups and supported ongoing classification initiatives by identifying vast classes of simple groups. A landmark result came in 1963 with the Feit-Thompson theorem, which proves that every finite group of odd is solvable, resolving a long-standing and eliminating odd-order nonsimple groups from consideration in classifications. The proof, spanning over 250 pages and employing intricate and formation theory, showed that no nonabelian simple group of odd order exists, thereby restricting potential s to even order. This theorem served as a cornerstone for the (CFSG), narrowing the scope of the project initiated in the . The CFSG, a monumental collaborative effort spanning the 1960s to 2004, culminated in the theorem that every finite simple group is either cyclic of prime order, an alternating group, a group of Lie type, or one of 26 sporadic groups. Key contributions included Michael Aschbacher's 1980s program on subsystems and signalizer functors, which streamlined proofs for groups with BN-pair structures, and revisions by Robert Guralnick addressing gaps in character-theoretic arguments. Richard Lyons and Ronald Solomon, building on Daniel Gorenstein's foundational work, produced a second-generation proof in the 1990s-2000s, reorganizing the classification into manageable cases and verifying completeness by 2004. This classification not only enumerated all simple building blocks of finite groups but also enabled applications in number theory and geometry. Computer-assisted methods gained prominence with the 1985 publication of the ATLAS of Finite Groups, which compiled detailed tables of maximal subgroups, character tables, and constructions for all sporadic simple groups and many Lie-type groups up to certain ranks. Authored by John H. Conway and collaborators, the ATLAS facilitated verification of CFSG components through computational checks on representations and fusion systems, bridging theoretical proofs with explicit data. Its resources proved invaluable for identifying outer automorphisms and resolving ambiguities in the classification.

References

  1. [1]
    [PDF] Finite and Sporadic Groups - MIT Mathematics
    Jun 14, 2023 · A finite group is one with only a finite number of elements. Definition 2.5. The order of a finite group, written |𝐺|, is the number of ...
  2. [2]
    None
    ### Summary of Finite Groups Section from Abstract Algebra Notes by R.J. Buehler
  3. [3]
    [PDF] Notes on finite group theory - Queen Mary University of London
    This starts from the definition of a group and includes subgroups and homomorphisms, examples of groups, group actions, Sylow's theorem, and composition series.Missing: authoritative sources
  4. [4]
    None
    ### Key Facts About Finite Groups and Classification of Finite Simple Groups
  5. [5]
    [PDF] Definition and Examples of Groups
    Definition 21.1. A group is a nonempty set G equipped with a binary operation ∗ : G×G → G satisfying the following axioms: ı(i) Closure: if a, b ∈ G, ...
  6. [6]
    Group -- from Wolfram MathWorld
    A group is a monoid each of whose elements is invertible. A group must contain at least one element, with the unique (up to isomorphism) single-element group ...
  7. [7]
    Finite Group -- from Wolfram MathWorld
    A finite group is a group having finite group order. Examples of finite groups are the modulo multiplication groups, point groups, cyclic groups, dihedral ...Missing: authoritative sources
  8. [8]
    finite group in nLab
    Sep 7, 2023 · Definition. A finite group is a group whose underlying set is finite. This is equivalently a group object in FinSet.Missing: authoritative sources
  9. [9]
    [PDF] ORDERS OF ELEMENTS IN A GROUP 1. Introduction Let G be a ...
    When G is a finite group, every element must have finite order. However, the converse is false: there are infinite groups where each element has finite order.
  10. [10]
    Order of element divides order of group - Groupprops
    Oct 25, 2008 · Specifically, the (order of an element) of a/an/the (element of a group) divides the (order of a group) of a/an/the (group). View other divisor ...
  11. [11]
    [PDF] cyclic-groups.pdf
    In terms of Zn, this result says that m ∈ Zn has order n (m, n) . Example. (Finding the order of an element) Find the order of the element a32 in the cyclic ...
  12. [12]
    [PDF] Order (group theory)
    All elements of finite groups have finite order. The order of a group G is denoted by ord(G) or |G| and the order of an element a by ord(a) or |a|. Example.
  13. [13]
    [PDF] Cosets and Lagrange's theorem - Keith Conrad
    In particular, when H is finite, the cosets of H all have the same size as H. Proof. Pick a left coset, say gH. We can pass from gH to H by left multiplication ...
  14. [14]
    [PDF] no subgroup of a4 has index 2 - Keith Conrad
    Page 2. 2. KEITH CONRAD. Theorem 4. If G is a finite group with a subgroup of index 2 then its commutator subgroup has even index. Proof. If [G : H] = 2 then H ...
  15. [15]
    [PDF] MAT 312/AMS 351 Notes and exercises on normal subgroups and ...
    A normal subgroup H of G means gH = Hg for every g in G. A quotient group G/H is formed when H is normal, where cosets form a group.
  16. [16]
    [PDF] proof of cauchy's theorem - Keith Conrad
    Theorem. (Cauchy) Let G be a finite group and p be a prime factor of |G|. Then G contains an element of order p. Equivalently, G contains a subgroup of order p.
  17. [17]
    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.
  18. [18]
    [PDF] EULER'S THEOREM 1. Introduction Fermat's little ... - Keith Conrad
    It is the relative primality point of view on the right that lets Fermat's little theorem be extended to a general modulus, as Euler discovered. Theorem 1.2 ( ...
  19. [19]
    [PDF] Math 403 Chapter 5 Permutation Groups: 1. Introduction
    A permutation group of A is a set of permutations of A that forms a group under function composition.
  20. [20]
    [PDF] 18.703 Modern Algebra, Permutation groups - MIT OpenCourseWare
    Definition 5.5. The group Sn is the set of permutations of the first n natural numbers. We want a convenient way to represent an element of Sn ...
  21. [21]
    [PDF] The Alternating Group
    The alternating group (An) is the set of even permutations in Sn, which is the kernel of δ, and is a normal subgroup of Sn.
  22. [22]
    [PDF] 1.6 Symmetric, Alternating, and Dihedral Groups
    The set An of all even permutations of Sn forms a subgroup of Sn of index 2. So An is a normal subgroup of Sn with |An| = n!/2. An is called the alternating ...Missing: A_n | Show results with:A_n
  23. [23]
    [PDF] SIMPLICITY OF An - KEITH CONRAD
    For n ≥ 5, the group An is simple. This is due to Camille Jordan [6, p. 66] in 1870. The special case n = 5 goes back to. Galois. The restriction n ≥ 5 is ...
  24. [24]
    [PDF] Permutation Groups
    Definition. A transposition is a permutation which interchanges two elements and leaves everything else fixed. (That is, a transposition is a cycle of length 2.).
  25. [25]
    Groups of small order. - UCSD Math
    Order 6 (2 groups: 1 abelian, 1 nonabelian). C_6; S_3, the symmetric group of degree 3 = all permutations on three objects, under composition. In cycle notation ...
  26. [26]
    [PDF] dihedral groups - keith conrad
    Introduction. For n ≥ 3, the dihedral group Dn is defined as the rigid motions1 taking a regular n-gon back to itself, with the operation being composition.Missing: symmetries | Show results with:symmetries
  27. [27]
    [PDF] Permutation Groups Rubik's Cube
    May 6, 2000 · In this paper we'll discuss permutations (rearrangements of objects), how to combine them, and how to construct complex permutations from ...
  28. [28]
  29. [29]
    None
    Below is a merged summary of cyclic groups based on the provided segments from "Group Theory" by James S. Milne (GT.pdf). To retain all information in a dense and organized manner, I will use a combination of narrative text and a table in CSV format for key details. The summary integrates definitions, structures, subgroups, and properties across all segments, ensuring no information is lost.
  30. [30]
    13.1: Finite Abelian Groups - Mathematics LibreTexts
    Jun 4, 2022 · We shall prove the Fundamental Theorem of Finite Abelian Groups which tells us that every finite abelian group is isomorphic to a direct product ...Missing: primary | Show results with:primary
  31. [31]
    Invariant Factor and Elementary Divisor Calculator | Ex Libris
    Jun 9, 2016 · The Fundamental Theorem of Finite Abelian Groups decisively characterizes the Abelian finite groups of a given order.<|control11|><|separator|>
  32. [32]
    [PDF] The Theory of p-Groups
    Definition 2.5 Let G be a finite p-group, of order pn. If c denotes the class of G, then the coclass of G is the quantity n − c. Having failed completely ...
  33. [33]
    [PDF] 3 p-groups - Brandeis
    One of the most important properties of p-groups is that they have nontrivial centers: Theorem 3.1. Every nontrivial p-group has a nontrivial center (P 6= 1 ⇒.
  34. [34]
    Sur certains groupes simples - Project Euclid
    1955 Sur certains groupes simples. C. Chevalley · DOWNLOAD PDF + SAVE TO MY LIBRARY. Tohoku Math. J. (2) 7(1-2): 14-66 (1955). DOI: 10.2748/tmj/1178245104.
  35. [35]
    [PDF] the sylow theorems - keith conrad
    In the proof of Sylow I, we saw that if H is a p-subgroup of G that is not a p-Sylow subgroup then N(H) is strictly larger than H. What can be said about N(P) ...
  36. [36]
    [PDF] Notes on the Proof of the Sylow Theorems
    Theorem 1.4 Any two Sylow p-subgroups are conjugate. Proof Let P be any Sylow p-subgroup. Let S0 be the set of all G-conjugates of P. Then S0 is.
  37. [37]
    [PDF] 23 Proofs and Applications of the Sylow Theorems
    1. Recall that a Sylow p-subgroup is a subgroup H ≤ G such that |H| = p . The first theorem states that there always exists a Sylow p-subgroup. 2.
  38. [38]
    [PDF] For a group theorist, Sylow's Theorem is such a basic tool, and so ...
    Characterizing Cyclic Groups. The Sylow theorems can be used to prove some theorems about all finite groups after they are proved for p-groups. Here is an ...
  39. [39]
    [PDF] 21. Sylow Theorems and applications - UCSD Math
    Sylow Theorems and applications. In general the problem of classifying groups of every order is com- pletely intractable. Given any group G, the first thing ...
  40. [40]
    [PDF] groups of order 12 - keith conrad
    Since each 3-Sylow subgroup has order 3, different 3-Sylow subgroups intersect trivially. Each of the 3-Sylow subgroups of G contains two elements of order 3, ...
  41. [41]
    [PDF] Finite Group Theory
    Burnside's normal p-complement theorem extends this result to all primes. Corollary 1.22 Let G be a finite group, and p the smallest prime dividing |G|. If the ...
  42. [42]
    VII. On the theory of groups, as depending on the symbolic equation ...
    On the theory of groups, as depending on the symbolic equationθ n =1. A. Cayley Esq. 2 Stone Buildings. Pages 40-47 | Published online: 30 Apr 2009.
  43. [43]
    [PDF] Proof of Cayley's Theorem, and an example: G = S3
    Proof of Cayley's Theorem, and an example: G = S3. 1. We define ψ from G to the set of all functions G → G by, for a in G, ψ(a) is left.
  44. [44]
    [PDF] Cayley's Theorem - Home | Department of Mathematics
    Sep 26, 2016 · Let G be a finite group of order n. Then G is isomorphic to a subgroup of. Sn, the symmetric group on n letters. Proof. Cayley's Theorem gives ...
  45. [45]
    [PDF] Cayley's Theorem
    The proof of this generalization is similar to that of Cayley's Theorem. We list the distinct cosets in G{H as a1H,...,anH. Then define ϕ ∶ G Ñ Sn by ϕpaq ...
  46. [46]
    [PDF] Theory of Groups of Finite Order - Project Gutenberg
    Title: Theory of Groups of Finite Order. Author: William Burnside. Release Date: August 2, 2012 [EBook #40395]. Language: English. Character set encoding: ISO ...
  47. [47]
    Burnside Lemma - Encyclopedia of Mathematics
    Dec 1, 2014 · [a1], W. Burnside, "Theory of groups of finite order" , Cambridge Univ. Press (1897) ; [a2], W. Burnside, "Theory of groups of finite order" , ...
  48. [48]
    14.3: Burnside's Counting Theorem - Mathematics LibreTexts
    Sep 29, 2021 · Burnside's The Theory of Groups of Finite Order, published in 1897, was one of the first books to treat groups in a modern context as ...
  49. [49]
    [PDF] Analysis and Applications of Burnside's Lemma - MIT Mathematics
    May 17, 2018 · Abstract. Burnside's Lemma, also referred to as Cauchy-Frobenius Theorem, is a result of group theory that is used to count distinct objects.
  50. [50]
    9.2: Direct Products - Mathematics LibreTexts
    Sep 7, 2021 · The external direct product of two groups builds a large group out of two smaller groups. We would like to be able to reverse this process and ...External Direct Products · Example 9.15 · Corollary 9.23 · Internal Direct Products
  51. [51]
    direct products of groups - PlanetMath
    Mar 22, 2013 · The direct product of two groups G G and H H is usually written G×H G × H , or sometimes G⊕H G ⊕ H (or G∐H G ⁢ ∐ H ) if G G and H H are both ...
  52. [52]
    [PDF] Direct Products
    For example, |Z5 × Z6| = 30. Lemma. The product of abelian groups is abelian: If G and H are abelian, so is G × H. Proof. Suppose G and H are abelian. Let ...
  53. [53]
    4.2: Product Groups - Mathematics LibreTexts
    Mar 5, 2022 · The direct product (or just product) of two groups G and H is the group G × H with elements ( g , h ) where g ∈ G and h ∈ H . The group ...
  54. [54]
    [PDF] SEMIDIRECT PRODUCTS 1. Introduction For two groups H and K ...
    For example, in S3 if H = h(12)i and K = h(13)i then HK = {(1),(12),(13),(132)} has size. 4 and is not a subgroup of S3. However, if H or K is normal in G ...
  55. [55]
    [PDF] Supplement. Direct Products and Semidirect Products
    Feb 5, 2018 · We also define the semidirect product and illustrate its use in classifying groups of small order. The content of this supplement is based on.
  56. [56]
    Solvable Group -- from Wolfram MathWorld
    A solvable group is a group having a normal series such that each normal factor is Abelian. The special case of a solvable finite group is a group whose ...
  57. [57]
    [PDF] SOLVABLE GROUPS Definition. Suppose that G is a finite group ...
    It can be proved that if G is a solvable group, then every subgroup of G is a solvable group and every quotient group of G is also a solvable group. Suppose ...
  58. [58]
    [PDF] Solvable group
    solvable group G. For finite groups, an equivalent definition is that a solvable group is a group with a composition series all of whose factors are cyclic ...
  59. [59]
    finite nilpotent groups - PlanetMath.org
    Mar 22, 2013 · 1. G G is nilpotent. · 2. Every Sylow subgroup of G G is normal. · 3. For every prime p∣∣|G| p | | G | , there exists a unique Sylow p p -subgroup ...
  60. [60]
    Burnside's normal p-complement theorem - Groupprops
    ### Statement of Burnside's Normal p-Complement Theorem
  61. [61]
    [PDF] Composition Series and Jordan-Holder
    May 11, 2024 · An important fact is that all finite groups have a composition series. The argument goes like this: suppose we know that all groups with ...
  62. [62]
    [PDF] Composition Series and the Hölder Program
    A composition series is a sequence of subgroups in a group. The Jordan-Hölder theorem states that a finite group has a composition series, and the factors are ...
  63. [63]
    [PDF] 5 Notes on Jordan-Hölder - Brandeis
    Theorem 5.4 (Schreier Refinement Theorem). Any two subnormal se- ries for G have equivalent refinements where equivalent means the sequences of subquotients ...
  64. [64]
    [PDF] SUBGROUP SERIES II 1. Introduction In part I, we met nilpotent and ...
    (1) every minimal normal subgroup has prime order and lies in the center,. (2) every maximal subgroup M is normal with prime index and contains the commutator.
  65. [65]
    [PDF] MATH 3030, Abstract Algebra FALL 2012 Toby Kenney Basic ...
    Compute a composition series for S4. We know that A4 is a normal subgroup of S4, so we can choose A4 as one el- ement in the composition series.
  66. [66]
    [PDF] Chapter 25 Finite Simple Groups
    Thus, set of elements in G corresponds to even permutations in SG is a normal subgroup of index 2. Thus, it is a normal subgroup of G. Chapter 25 Finite ...
  67. [67]
    A Simple Abelian Group if and only if the Order is a Prime Number
    We prove that a group is an abelian simple group if and only if the order of the group is prime number. Any group of prime order is a cyclic group, and abelian.Missing: source | Show results with:source
  68. [68]
    [PDF] On the Classification of Finite Simple Groups - MIT Mathematics
    May 22, 2022 · In this paper, we explore the properties and applications of finite simple groups. We begin by discussing basic definitions and theorems in ...
  69. [69]
    [PDF] Supplement. Finite Simple Groups
    Recall that a group is simple if it is nontrivial and has no proper nontrivial normal subgroups. (Definition 15.14 of Fraleigh). Note. Since the cyclic groups ...
  70. [70]
    OUTER AUTOMORPHISMS OF GROUPS - Project Euclid
    Jul 29, 1988 · It is obvious that if G is a group embedded in another group H, then any inner automorphism of G lifts to an (inner) automorphism of H. Angus ...
  71. [71]
    SCHUR MULTIPLIERS OF THE KNOWN FINITE SIMPLE GROUPS1 ...
    ABSTRACT. In this note, we announce some results about the Schur multipliers of the known finite simple groups. Proofs will appear else- where.
  72. [72]
    [PDF] The Classification of Finite Simple Groups An enormous theorem
    The classification of finite simple groups, completed in 1981, is an extraordinary theorem that aimed to classify all simple groups, which are building blocks ...
  73. [73]
    Sporadic Group -- from Wolfram MathWorld
    The sporadic groups are the 26 finite simple groups that do not fit into any of the four infinite families of finite simple groups.
  74. [74]
    Enumeration of Finite Groups - Cambridge University Press
    Enumeration of Finite Groups ... Blackburn, Royal Holloway, University of London, Peter M. Neumann, University of Oxford, Geetha Venkataraman, University of Delhi.
  75. [75]
    Enumerating Finite Groups of Given Order - jstor
    Let f(n) denote the number of (isomorphism classes of) groups of order n. Let n = HW p9ig be the prime factorization of n and let A = A(n) = Ek and ,u = ii ...
  76. [76]
    [PDF] ENUMERATING P-GROUPS BETTINA EICK and E. A. O'BRIEN ...
    Sep 6, 1999 · Higman (1960) and Sims (1965) provide asymptotic esti- mates which show that the number of groups of order pn is p2nõ/27+O(nûüõ). The p-group ...
  77. [77]
    GAP (ref) - Chapter 50: Group Libraries
    There are several infinite families of groups which are parametrized by numbers. GAP provides various functions to construct these groups.
  78. [78]
    [2501.09769] Classifying the groups of order $p q$ in Lean - arXiv
    Jan 16, 2025 · This note discusses our formalisation in Lean of the classification of the groups of order pq for (not necessarily distinct) prime numbers p and q.
  79. [79]
    Joseph-Louis Lagrange (1736 - 1813) - Biography - MacTutor
    He studied permutations of the roots and, although he does not compose permutations in the paper, it can be considered as a first step in the development of ...
  80. [80]
    Paolo Ruffini - Biography
    ### Summary of Ruffini's Work on Solvability by Radicals in 1799
  81. [81]
    Augustin-Louis Cauchy (1789 - 1857) - Biography - MacTutor
    Augustin-Louis Cauchy pioneered the study of analysis, both real and complex, and the theory of permutation groups. He also researched in convergence and ...
  82. [82]
    Évariste Galois - Biography
    ### Summary of Galois's Work in the 1830s on Galois Groups and Solvability by Radicals
  83. [83]
    [PDF] The mathematical writings of Evariste Galois - Uberty
    An article published in June 1830 created the theory of Galois imaginaries, a fore-runner of what are now known as finite fields; his so-called Premier ...
  84. [84]
    On the Theory of Groups - jstor
    On the Theory of Groups. By PROF. CAYLEY. I refer to my papers on the theory of groups as depending on the symbolic.
  85. [85]
    Arthur Cayley and the First Paper on Group Theory (Chapter 1)
    Summary. Arthur Cayley's 1854 paper On the theory of groups, as depending on the symbolic equation θn = 1 inaugurated the abstract idea of a group [2].
  86. [86]
    Traité des substitutions et des équations algébriques - Internet Archive
    Dec 3, 2008 · This book, 'Traité des substitutions et des équations algébriques', by Camille Jordan, published in 1870, covers topics like Galois theory and ...
  87. [87]
    [PDF] 36154-pdf.pdf - Project Gutenberg
    The theory of differential equations is just now being cultivated very extensively by French mathematicians; and some of them proceed precisely from this point ...
  88. [88]
    Felix Klein's projective representations of the groups $$S_6$$ and ...
    Apr 25, 2022 · This paper addresses an article by Felix Klein of 1886, in which he generalized his theory of polynomial equations of degree 5.
  89. [89]
    Théorèmes sur les groupes de substitutions - EuDML
    Sylow. "Théorèmes sur les groupes de substitutions." Mathematische Annalen 5 (1872): 584-594. <http://eudml.org/doc/156588>.
  90. [90]
    Théorèmes sur les groupes de substitutions | Mathematische Annalen
    Sylow, M.L. Théorèmes sur les groupes de substitutions. Math. Ann. 5, 584–594 (1872). https://doi.org/10.1007/BF01442913. Download citation. Issue Date: ...Missing: URL | Show results with:URL
  91. [91]
    A history of the Burnside problem - MacTutor - University of St Andrews
    A history of the Burnside problem ... It is clear that any finite group is periodic. In his 1902 paper, Burnside [1] introduced what he termed "a still ...
  92. [92]
    [PDF] On the Burnside Problem - UChicago Math
    Aug 12, 2006 · This paper will investigate the classical Burnside problem as it was originally proposed in 1902 and will focus on developing a counterexample.
  93. [93]
    ON BURNSIDE'S OTHER paqh THEOREM - Project Euclid
    Suppose G is a finite group whose order is divisible by only two primes. Burnside's famous theorem asserts that G must be solvable. In a less famous theorem ...
  94. [94]
    [PDF] Theory of groups of finite order / W. Burnside.
    Theory of groups of finite order / W. Burnside. Burnside, William, 1852-1927. Cambridge: The University press, 1911. http://hdl.<|separator|>
  95. [95]
    Issai Schur (1875 - 1941) - Biography - MacTutor
    Schur is mainly known for his fundamental work on the representation theory of groups but he also worked in number theory, analysis and other topics.Missing: finite | Show results with:finite
  96. [96]
    [PDF] Contributions of Issai Schur to Analysis - arXiv
    Jun 13, 2007 · The name Schur is associated with many terms and concepts that are widely used in a number of diverse fields of mathematics and engineering.
  97. [97]
    Claude Chevalley (1909 - 1984) - Biography - MacTutor
    [The Theory of Lie Groups] was the first systematic exposition of the foundations of Lie group theory consistently adopting the global viewpoint, based on the ...
  98. [98]
    Chevalley Groups -- from Wolfram MathWorld
    The Chevalley groups are the finite simple groups of Lie-type. They include four families of linear simple groups.
  99. [99]
    Chapter I, from Solvability of groups of odd order, Pacific J ... - MSP
    THEOREM. All finite groups of odd order are solvable. Some consequences of this theorem and a discussion of the proof may be found in ...
  100. [100]
    Volume 13 Issue 3 | Pacific Journal of Mathematics - Project Euclid
    Chapter I, from Solvability of groups of odd order, Pacific J. Math, vol. 13, no. 3 (1963) Walter Feit, John G. Thompson
  101. [101]
    Feit-Thompson Theorem -- from Wolfram MathWorld
    Every finite simple group (that is not cyclic) has even group order, and the group order of every finite simple noncommutative group is doubly even.
  102. [102]
    The Classification of Finite Simple Groups
    This book outlines the proof of the Classification of Finite Simple Groups, focusing on groups of characteristic 2 type, as the second half of the project.
  103. [103]
    [PDF] CFSG—A User's Manual: Applying the Classification of Finite Simple ...
    Aug 25, 2015 · As one reference on the CFSG, I often use the “outline”: • Aschbacher–Lyons–Smith–Solomon [ALSS11]. ... Aschbacher-Guralnick [AG84, Thm B] showed ...
  104. [104]
    Atlas of Finite Groups - John Horton Conway - Oxford University Press
    This Atlas brings together detailed information about these groups--their construction, character tables, maximal subgroups, and more.Missing: 1985 | Show results with:1985
  105. [105]
    The Atlas of Finite Groups - Ten Years On
    The Atlas of Finite Groups by J. H. Conway, R. A. Parker, S. P. Norton and the editors of this book, was published in 1985, and has proved itself to be an ...
  106. [106]
    [PDF] FINITE GROUPS - Gaetan Chenevier
    ... Atlas of finite groups. Bibliography: p. Indudes index. 1. Finite groups. I. Conway, John Horton. ()JlI71.Jl86 1985. 512'.2. 85-11559. ISBN 0-19-853199-0.