Fact-checked by Grok 2 weeks ago

Permutation group

In , particularly within , a permutation group is a group whose elements are permutations of a given and whose binary operation is the of such permutations, forming a subgroup of the symmetric group on that set. The symmetric group S_n, which consists of all possible permutations of a set with n elements, serves as the foundational example and has order n!, representing the total number of ways to rearrange the elements. A key result in group theory, known as Cayley's theorem, establishes that every is isomorphic to a permutation group acting regularly on itself, thereby embedding abstract groups into the concrete framework of permutations. This theorem underscores the universality of permutation groups in modeling symmetries and structures across . Permutations within these groups can be classified by their —even or odd—based on the parity of the number of transpositions needed to express them, leading to the A_n, the kernel of the sign homomorphism and a of S_n of index 2. For n \geq 5, A_n is a , meaning it has no nontrivial normal subgroups, which has profound implications for the . Permutation groups play a central role in various applications, including the study of geometric symmetries, where they describe the automorphism groups of polytopes and graphs; in , for enumerating objects under group actions via tools like ; and in physics and chemistry, for analyzing molecular symmetries and identical particle systems. In computational contexts, algorithms for permutation group manipulation, such as membership testing and finding strong generators, enable efficient solutions to problems in and . These structures also connect to , where permutation groups of roots illuminate the solvability of equations.

Fundamentals

Definition

In group theory, a permutation group G on a set X is defined as a of the \Sym(X), where the elements of G are bijections from X to itself, and the group operation is the of these bijections. The \Sym(X) comprises all permutations of the set X, that is, all bijective functions from X to X, equipped with the operation of , which forms a group under this operation. When X is finite with cardinality n, the order of \Sym(X) is n!, reflecting the number of distinct bijections possible. Permutation groups may be finite or infinite, with the infinite case arising when the underlying set X is infinite, leading to an uncountably many permutations in \Sym(X) for uncountable X. A basic example is the trivial group, which serves as the permutation group on a singleton set X = \{ a \}, where \Sym(X) consists solely of the identity bijection that maps a to itself.

Basic Properties

A permutation group G on a set X satisfies the closure property under the group operation of composition: the composition of any two elements \sigma, \tau \in G is another bijection from X to X, hence also in G. This ensures that G forms a subgroup of the symmetric group \mathrm{Sym}(X), the full group of all bijections on X. The operation of composition in a permutation group inherits associativity from the associativity of : for any \sigma, \tau, \rho \in G, (\sigma \circ \tau) \circ \rho = \sigma \circ (\tau \circ \rho). When X is finite with |X| = n, any permutation group G \leq \mathrm{Sym}(X) is finite, and by applied to the finite group \mathrm{Sym}(X) of order n!, the order of G divides n!. Every in a permutation group admits a unique into a product of disjoint , up to ordering of the and within each ; this reveals the of the by partitioning the moved points into orbits under repeated application. The cycle type of a is the of lengths of these disjoint , providing a complete for its on X. The support of a is the subset of X consisting of points not fixed by the , i.e., those appearing in its (excluding 1-cycles). Two permutations in \mathrm{Sym}(X) are conjugate if and only if they have the same cycle type; that is, \tau^{-1} \sigma \tau has the same cycle structure as \sigma for some \tau \in \mathrm{Sym}(X). In a G \leq \mathrm{Sym}(X), permutations with the same cycle type belong to the same in \mathrm{Sym}(X) but may form multiple conjugacy classes in G.

Notation and Operations

Notation Conventions

Permutations of a , such as \{1, 2, \dots, n\}, are often represented in one-line notation as \sigma = (\sigma(1) \ \sigma(2) \ \dots \ \sigma(n)), where the sequence lists the images of the elements under the permutation \sigma. This compact form emphasizes the functional mapping directly. Another common representation is the two-row notation, written as \begin{pmatrix} 1 & 2 & \dots & n \\ \sigma(1) & \sigma(2) & \dots & \sigma(n) \end{pmatrix}, which explicitly pairs each domain element with its image. This format highlights the between the top and bottom rows. Cycle notation decomposes a into its disjoint , such as (1 \ 3 \ 2)(4 \ 5), where each indicates a cyclic shift and fixed points are typically omitted. are written with elements in sequence, starting from an arbitrary point, and the notation is unique up to and ordering of cycles. In permutation group actions, the convention often distinguishes left and right actions; a right action applies the permutation post-fix, denoted x^\sigma for x in the set acted upon, which aligns with standard algebraic composition where permutations multiply on the right. This right-action preference avoids reversal in group multiplication orders, unlike the left-action form \sigma(x). For imprimitive actions, blocks of imprimitivity are denoted as subsets B_1, B_2, \dots, B_k forming a of the set, where the group permutes these blocks as units while preserving the . Such notation captures the non-trivial classes under the .

Composition as Group Product

In permutation groups, the group operation is defined as the of , which are bijective functions on a . For two permutations \sigma and \tau, their product \sigma \tau is the permutation that applies \tau first and then \sigma, formally given by (\sigma \tau)(x) = \sigma(\tau(x)) for all x in the set. This convention aligns with the standard right-to-left order of , where the rightmost permutation is applied initially in any sequence. When computing the product of permutations expressed in cycle notation, the mappings are evaluated sequentially from right to left for each element in the set. To determine the image of an element under the product, start by applying the rightmost , then proceed leftward through the cycles, tracking the position until returning to the starting element or fixing it. For instance, consider the product (1\ 2)(2\ 3): applying (2\ 3) followed by (1\ 2) yields $1 \mapsto 2 \mapsto 3, $3 \mapsto 2 \mapsto 1, and $2 \mapsto 3 \mapsto 3 (fixed after the first step, but the cycle closes as (1\ 2\ 3)); thus, (1\ 2)(2\ 3) = (1\ 2\ 3). This process ensures the resulting is expressed in disjoint cycle form by identifying all cycles in the mapping. The composition operation is associative, as permutations are functions and function composition satisfies (\sigma \tau) \rho = \sigma (\tau \rho) for any permutations \sigma, \tau, \rho. To see this, evaluate both sides on an arbitrary x: the left side gives \sigma(\tau(\rho(x))), while the right side gives \sigma((\tau \rho)(x)) = \sigma(\tau(\rho(x))), confirming equality. This associativity underpins the group structure. The product of any two elements in a permutation group yields another element within the group, ensuring under . This property generates new permutations while preserving the group's finite order, as all elements are bijections on a set of fixed size, and repeated compositions remain within the set of all possible permutations.

Identity and Inverses

In permutation groups, the is the permutation that maps every of the to itself, denoted by e or id, and formally defined as id(x) = x for all x in the set./15:_Group_Theory_and_Applications/15.03:_Permutation_Groups) This serves as the neutral element under , satisfying \sigma \circ id = id \circ \sigma = \sigma for any permutation \sigma. Every permutation \sigma has a unique \sigma^{-1} such that \sigma \circ \sigma^{-1} = \sigma^{-1} \circ \sigma = id./15:_Group_Theory_and_Applications/15.03:_Permutation_Groups) Since permutations are s—both injective and surjective—they are invertible in the set of all functions from the to itself, with the inverse also being a bijection. To compute the inverse in cycle notation, reverse the order of elements in each cycle while preserving the disjoint cycle structure; for example, the inverse of (1\ 2\ 3) is (3\ 2\ 1), which is equivalent to (1\ 3\ 2)./15:_Group_Theory_and_Applications/15.03:_Permutation_Groups) The order of a permutation \sigma, denoted |\sigma|, is the smallest positive integer k such that \sigma^k = id. For a permutation expressed as a product of disjoint cycles of lengths l_1, l_2, \dots, l_m, the order is the of these lengths, |\sigma| = \operatorname{lcm}(l_1, l_2, \dots, l_m)./15:_Group_Theory_and_Applications/15.03:_Permutation_Groups) This follows from the fact that powers of disjoint cycles act independently, and each cycle of length l_i returns to the after l_i applications.

Examples

Symmetric and Alternating Groups

The S_n consists of all permutations of a set with n elements, which can be viewed as the set of all bijections from \{1, 2, \dots, n\} to itself under . The order of S_n is n!, reflecting the number of distinct bijections on an n-element set. Furthermore, S_n is generated by the set of all transpositions, the permutations that swap exactly two elements and fix the rest, for n \geq 2. The A_n is the of S_n comprising all even permutations, defined as those that can be expressed as a product of an even number of transpositions. For n \geq 2, A_n has 2 in S_n, meaning |A_n| = n!/2, and it is a . Additionally, A_n is simple for n \geq 5, possessing no nontrivial subgroups. The distinction between even and odd permutations arises from the sign homomorphism \operatorname{sgn}: S_n \to \{ \pm 1 \}, which assigns +1 to even permutations and -1 to odd ones, based on the parity of the number of transpositions in their decomposition. The kernel of this homomorphism is precisely A_n. A concrete example is S_3, the on three elements, which has order 6 and includes all permutations such as the , transpositions like (1\ 2), and 3-cycles like (1\ 2\ 3). The A_3 consists of the even permutations: the and the two 3-cycles (1\ 2\ 3) and (1\ 3\ 2), forming a of order 3 generated by (1\ 2\ 3).

Cyclic and Dihedral Groups

The C_n of order n provides a fundamental example of a permutation group, realized as the generated by the n- \sigma = (1\ 2\ \dots\ n) acting on the set \{1, 2, \dots, n\}. The elements of C_n are the powers \sigma^k for k = 0, 1, \dots, n-1, with \sigma^0 being the identity permutation, and the group operation is composition of permutations. Since powers of a single generator commute under composition, C_n is abelian. A key extension of the in the context of permutation groups is the D_n (also denoted D_{2n} in some conventions), which captures the full of a regular n-gon and has order $2n. This group acts faithfully on the n vertices of the polygon as and is generated by a \rho = (1\ 2\ \dots\ n), which cycles the vertices, and a s, which swaps pairs of vertices symmetrically while fixing one (for odd n) or passing through midpoints of edges (for even n). The defining relations are \rho^n = e, s^2 = e, and s \rho s = \rho^{-1}, reflecting how reflections conjugate rotations to their inverses. The abstract presentation of D_n is \langle \rho, s \mid \rho^n = s^2 = 1, s \rho s^{-1} = \rho^{-1} \rangle, which encodes its structure as a of the C_n by C_2. For n = 3, the D_3 consists of the three rotations and three reflections of an , permuting its vertices, and is isomorphic to the S_3 of order 6.

Group Actions

Definition and Examples

In the context of permutation groups, a provides a way to realize the elements of a group G as of a set X. Formally, a left action of G on X is a \phi: G \to \Sym(X), where \Sym(X) denotes the consisting of all bijections from X to itself under . Equivalently, it can be described by a map G \times X \to X, written (g, x) \mapsto g \cdot x, that satisfies two axioms: the e \in G acts trivially, so e \cdot x = x for all x \in X, and the action is compatible with the group operation, so (gh) \cdot x = g \cdot (h \cdot x) for all g, h \in G and x \in X. This homomorphism perspective highlights that every induces a of G, namely the image \phi(G), which is a of \Sym(X) acting naturally on X. The permutation representation arising from a group action captures how G permutes the elements of X; distinct elements of G may induce the same permutation if the action is not injective. A key distinction is between faithful and non-faithful actions: an action is faithful if the kernel of \phi is trivial, meaning \ker \phi = \{ e \}, so that only the identity element fixes every point in X, and thus \phi embeds G as a subgroup of \Sym(X). Otherwise, the action is non-faithful, and the kernel is a normal subgroup of G consisting of all elements that act as the identity permutation. Illustrative examples clarify these concepts. The trivial action of any group G on a set X is defined by g \cdot x = x for all g \in G and x \in X; this always yields a valid , but it is faithful only if G is the , and the corresponding permutation is the singleton subgroup of \Sym(X). A more substantive example is the left regular , where G acts on itself as the set X = G by left multiplication: g \cdot h = gh for g, h \in G. This is always faithful, producing a permutation that embeds G isomorphically into \Sym(G), the full on |G| elements.

Orbits and Stabilizers

In group theory, when a permutation group G acts on a set X, the orbit of an element x \in X is defined as the set \operatorname{Orb}_G(x) = \{ g \cdot x \mid g \in G \}, where g \cdot x denotes the action of g on x. The orbits of all elements in X form a of X, meaning every element belongs to exactly one orbit, and distinct orbits are disjoint. The stabilizer of x \in X, denoted \operatorname{Stab}_G(x) = \{ g \in G \mid g \cdot x = x \}, is the of G consisting of all permutations that fix x. This captures the symmetries preserved at the point x under the action. A fundamental relation between orbits and stabilizers is given by the orbit-stabilizer theorem. For a finite group G acting on X, the size of the of x equals the of the in G: |\operatorname{Orb}_G(x)| = [G : \operatorname{Stab}_G(x)] = \frac{|G|}{|\operatorname{Stab}_G(x)|}. This theorem arises from the between the cosets of \operatorname{Stab}_G(x) in G and the elements of \operatorname{Orb}_G(x), where the coset g \operatorname{Stab}_G(x) maps to g \cdot x. The core of the stabilizer \operatorname{Stab}_G(x) is the largest of G contained in \operatorname{Stab}_G(x), defined as \operatorname{Core}_G(\operatorname{Stab}_G(x)) = \bigcap_{g \in G} g \operatorname{Stab}_G(x) g^{-1}, the of all conjugates of the . This core measures the "essential" normality properties embedded in the pointwise symmetries. For example, consider the S_3 acting naturally on the set \{1, 2, 3\} by permuting its elements. The orbit of 1 is \operatorname{Orb}_{S_3}(1) = \{1, 2, 3\}, with |\operatorname{Orb}_{S_3}(1)| = 3, while the \operatorname{Stab}_{S_3}(1) = \{\operatorname{id}\}, the trivial of 1. The confirms $3 = 6 / 1, and the core of the stabilizer is trivial since it is already normal (being trivial).

Transitive and Primitive Groups

Transitive Actions

A permutation group G on a X is said to be transitive if there is only one , meaning that for any two elements x, y \in X, there exists an element g \in G such that g \cdot x = y. This condition is equivalent to the action being such that G can map any point in X to any other point via some group element. The degree of a permutation group G acting on X is defined as the cardinality |X|. The minimal degree of G is the smallest possible degree of a transitive of G on a . For example, the S_n acts transitively on the set \{1, 2, \dots, n\} by its natural , which has degree n. Similarly, the A_n acts transitively on the same set for n \geq 3, as any even can map any element to any other while preserving the . A transitive action is doubly transitive if, for any two distinct ordered pairs (x, y) and (x', y') in X \times X with x \neq y and x' \neq y', there exists g \in G such that g \cdot x = x' and g \cdot y = y'. The S_n provides a classic example of a doubly transitive group on \{1, 2, \dots, n\} for n \geq 2, since any can rearrange distinct points arbitrarily. In contrast, A_n is 2-transitive only for n \geq 4 in certain contexts, but its basic holds more broadly. Transitive permutation groups may exhibit imprimitivity, characterized by the existence of blocks of imprimitivity. A nonempty proper B \subset X is a block of imprimitivity for G if, for every g \in G, either g \cdot B = B or g \cdot B \cap B = \emptyset. Such blocks form partitions of X that are preserved by the , indicating a coarser structure within the transitive action; for instance, the acting on the vertices of a may preserve subsets corresponding to opposite vertices as blocks.

Primitive Actions

In group theory, a permutation group G acting transitively on a \Omega with |\Omega| > 1 is if it preserves no nontrivial of \Omega. A \Delta \subseteq \Omega with $1 < |\Delta| < |\Omega| is a block of imprimitivity for G if, for every g \in G, either g\Delta = \Delta or g\Delta \cap \Delta = \emptyset; thus, primitivity means no such blocks exist. This condition distinguishes primitive actions from imprimitive ones, where nontrivial blocks allow the action to factor into coarser structures, such as when G acts on blocks and within blocks separately. An equivalent characterization is that the action is primitive if and only if the stabilizer G_\alpha of any point \alpha \in \Omega is a maximal subgroup of G. In this view, primitivity reflects the irreducibility of the action in terms of subgroup lattice: there are no intermediate subgroups between G_\alpha and G that could correspond to block systems. For instance, the symmetric group S_n acts primitively on the set \{1, \dots, n\} for n \geq 2, as the point stabilizer S_{n-1} is maximal. Similarly, the alternating group A_n acts primitively on \{1, \dots, n\} for n \geq 3, with point stabilizer A_{n-1} maximal. Imprimitive actions often arise from wreath products, which embed transitive but block-preserving groups. For example, the imprimitive wreath product S_k \wr S_m acts on a set of size km by permuting m blocks of size k, where the blocks form a nontrivial system of imprimitivity preserved by the group. This construction illustrates how imprimitivity permits decomposition into a product on blocks and a base group within blocks, contrasting with the "indecomposable" nature of primitive actions. The structure of finite primitive permutation groups is classified by the O'Nan-Scott theorem, which asymptotically divides them into five basic types based on the socle (minimal ): affine (socle elementary abelian, acting regularly), almost simple (socle nonabelian ), diagonal (socle of groups, acting by diagonal embedding), product (socle product of groups in action), and twisted wreath (socle in a twisted extension). This classification, refined over decades, underpins much of modern permutation group theory by linking primitivity to underlying structures without exhaustive enumeration.

Key Theorems

Cayley's Theorem

asserts that every group G is isomorphic to a of the \mathrm{Sym}(G) on the underlying set of G. This embedding arises from the left regular action of G on itself, defined by g \cdot h = gh for all g, h \in G. To prove this, consider the map \phi: G \to \mathrm{Sym}(G) given by \phi(g)(h) = gh for all h \in G. This defines a permutation of G for each g, as left multiplication by g is bijective: it has inverse left multiplication by g^{-1}. The map \phi is a because \phi(g_1 g_2)(h) = g_1 g_2 h = \phi(g_1)(\phi(g_2)(h)) for all g_1, g_2, h \in G. Moreover, \phi has trivial : if \phi(g) is the identity permutation, then g e = e, so g = e. Thus, \phi is injective, establishing the isomorphism G \cong \phi(G) \leq \mathrm{Sym}(G). The corresponding action is the regular action of G on itself by left , which is faithful (as \phi is injective) and transitive: for any h_1, h_2 \in G, there exists g = h_1 h_2^{-1} such that g \cdot h_2 = h_1. The degree of this permutation representation is |G|, and the stabilizer of any point h \in G is trivial, since \mathrm{Stab}(h) = \{g \in G \mid gh = h\} = \{e\}. As consequences, every group admits a faithful representation, allowing groups to be realized concretely as subgroups of symmetric groups. This shows that the class of groups is , encompassing all finite groups up to . For example, consider the C_3 = \langle r \mid r^3 = e \rangle. The regular action embeds C_3 into \mathrm{Sym}(C_3) \cong S_3 as the subgroup \{e, (1\ 2\ 3), (1\ 3\ 2)\}, corresponding to the rotations of an .

Structure of Permutation Groups

A permutation group, as a , is solvable if and only if its derived series terminates at the trivial after a finite number of steps, with the number of steps known as the derived length. The derived G' of a group G is generated by all commutators [g,h] = g^{-1}h^{-1}gh for g,h \in G, and the derived series is defined iteratively as G^{(0)} = G, G^{(k+1)} = (G^{(k)})'. Examples of solvable permutation groups include affine groups, such as the affine \mathrm{AGL}(1,p) for a prime p, which acts on p points as the \mathbb{Z}/p\mathbb{Z} \rtimes (\mathbb{Z}/p\mathbb{Z})^\times and has derived length 2 since both factors are abelian. Every finite permutation group admits a , a maximal chain of normal subgroups where each factor is , and by the Jordan-Hölder , the composition factors are unique up to and . The composition factors of permutation groups are thus non-abelian groups or cyclic groups of prime , with alternating groups A_k for k \geq 5 and cyclic groups frequently appearing due to the embedding of abstract groups into the via . For the S_n with n \geq 5, a is \{e\} \trianglelefteq A_n \trianglelefteq S_n, yielding composition factors A_n ( non-abelian) and \mathbb{Z}/2\mathbb{Z} (cyclic of prime ). The Frobenius theorem provides key insight into the structure of certain permutation groups possessing semiregular s. A N \trianglelefteq G acting semiregularly on a set means that only the of N has fixed points, implying trivial point stabilizers. The theorem states that if G is a finite permutation group with a nontrivial proper semiregular N, then N admits a complement H \leq G such that G = N \rtimes H, H \cap N = \{e\}, and H acts faithfully and semiregularly on the non-identity elements of N; such groups are precisely the Frobenius groups. This semidirect product decomposition reveals the internal structure, with N as the Frobenius kernel (regular if the action is transitive) and H as the Frobenius complement. Computational methods elucidate the structure of permutation groups given by generating sets. The Schreier-Sims algorithm constructs a and strong generating set (BSGS) for a group G \leq S_n, enabling efficient of the group via the product of lengths on the points and testing membership of elements in polynomial time relative to n \log |G|. It proceeds by building stabilizers iteratively: starting with a B = (b_1, \dots, b_k) such that the pointwise stabilizer G = G^{(0)} \geq G^{(1)} \geq \cdots \geq G^{(k)} = \{e\} has known orbits, and using Schreier generators from coset transversals to ensure strong generation. Extensions of the Schreier-Sims algorithm apply to black-box groups, where G is accessed via an for multiplication, inversion, and queries, allowing recognition and structure determination without explicit representation, as in nearly linear-time algorithms for .

Advanced Topics

Oligomorphic Groups

A permutation group G \leq \Sym(\Omega), where \Omega is typically a countably , is called oligomorphic if, for each k \geq 1, the action of G on the set \Omega^k of ordered k-tuples has only finitely many orbits. This finiteness condition contrasts with finite permutation groups, emphasizing bounded complexity in domains despite the group's potential largeness. The concept connects to through the age of a structure, defined as the of all finite substructures up to , which forms a under embeddings for relational structures with oligomorphic groups. A structure is homogeneous if every between its finite substructures extends to an of the whole ; such homogeneity often characterizes Fraïssé limits, where the is oligomorphic. Prominent examples include the \Aut(\mathbb{Q}, <) of the rational numbers under the strict order, which is oligomorphic and has exactly two orbits on the set of ordered pairs of distinct elements: one where the first is less than the second, and one where the order is reversed. Another is the of the (or random graph), the Fraïssé limit of the class of finite graphs, which is oligomorphic with the number of orbits on n-tuples equal to the number of non-isomorphic labeled graphs on n vertices up to adjacency relations. Countable oligomorphic permutation groups arise precisely as the automorphism groups of Fraïssé limits of their ages, as established in classifications of homogeneous structures; for instance, Lachlan and Woodrow's work on countable homogeneous graphs identifies the and certain Henson graphs as such limits with oligomorphic groups. These groups correspond to ω-categorical theories in , where the finiteness of orbits equates to the finiteness of n-types, enabling structural classifications. Oligomorphic groups find applications in for analyzing countable structures via their groups and in problems (CSPs), where structures with oligomorphic automorphisms allow complexity dichotomies; for example, valued CSPs over such templates reduce tractability to the existence of fractional polymorphisms or pp-constructions.

Permutation Group Isomorphisms

In group theory, two permutation groups G and H, where G acts faithfully on a set X and H acts faithfully on a set Y, are isomorphic as permutation groups if there exists a bijective \phi: G \to H and a \psi: X \to Y such that \phi(g) \cdot \psi(x) = \psi(g \cdot x) for all g \in G and x \in X. This condition ensures that the actions are preserved up to relabeling of the underlying sets, making the representations equivalent. If X = Y, the bijection \psi can be taken as the , reducing to the case where \phi is an of G that commutes with the action. Within the symmetric group \Sym(X), conjugacy provides a key notion of equivalence for elements and subgroups. Two permutations g, h \in \Sym(X) are conjugate if there exists \sigma \in \Sym(X) such that h = \sigma^{-1} g \sigma, denoted h = g^\sigma. This operation relabels the points of X via \sigma, preserving the cycle type of the permutation, which determines the . For subgroups, two subgroups H, K \leq \Sym(X) are conjugate if K = \sigma^{-1} H \sigma for some \sigma \in \Sym(X), implying they are isomorphic as abstract groups and their actions on X are equivalent up to relabeling. in \Sym(X) thus classify permutation representations up to such equivalences, with cycle types serving as invariants that distinguish non-conjugate elements. Equivalent representations of a group G on sets X and Y arise when there is a bijection \psi: X \to Y such that the image subgroups in \Sym(X) and \Sym(Y) are conjugate via \psi, meaning the actions are isomorphic. This equivalence captures how different embeddings of G into symmetric groups can represent the same underlying structure after adjusting for point labels. Burnside's lemma relates to these concepts by providing a tool to count orbits under group actions, which helps analyze the structure of permutation groups up to isomorphism. Specifically, for a group G acting on X, the number of orbits is given by |X/G| = \frac{1}{|G|} \sum_{g \in G} \Fix(g), where \Fix(g) is the number of fixed points of g. This formula, originally due to Frobenius and popularized by Burnside, quantifies the distinct "behaviors" of the action, aiding in distinguishing non-isomorphic representations. A example illustrates conjugacy in transitive actions: all transitive embeddings of the C_p of prime order p into \Sym(p) are conjugate. Here, each such embedding corresponds to the regular action of C_p on itself, and any two are related by conjugation in \Sym(p), as the transitive cyclic subgroups of prime degree are uniquely determined up to relabeling. This uniformity underscores how conjugacy normalizes representations, ensuring that isomorphic groups share equivalent transitive structures.

Historical Development

Early Contributions

The study of permutation groups originated in the late 18th century amid efforts to understand the solvability of polynomial equations. In his 1770 memoir Réflexions sur la résolution algébrique des équations, Joseph-Louis Lagrange examined permutations of equation roots as a means to analyze the structure of solutions for cubic and quartic equations, laying groundwork for linking algebraic resolvents to rearrangements of variables. Lagrange introduced rudimentary concepts of cycles by considering how repeated substitutions could generate solution functions, though he did not formalize them as group elements. Building on Lagrange's ideas, advanced the connection between s and equation solvability in his 1799 memoir Teoria generale delle equazioni, where he attempted to prove the impossibility of solving the general quintic equation by radicals using properties of arrangements of . Ruffini's argument, though incomplete, highlighted the role of the full in obstructing radical solutions for degree five polynomials. In 1824, provided a rigorous proof of the quintic's unsolvability, employing -based analysis to show that certain root rearrangements prevent expression via nested radicals, solidifying the permutation group's centrality to algebraic insolubility. Évariste Galois established the foundational framework for permutation groups in the 1830s through his work on equation solvability. In his 1831 Mémoire sur les conditions de résolubilité des équations par radicaux, Galois defined groups as closed sets of permutations under composition and linked solvability by radicals to the existence of a normal series with abelian factors in the symmetric group acting on roots. He demonstrated that the symmetric group S_n for n \geq 5 lacks such a series, rendering general polynomials of degree five or higher unsolvable by radicals. A pivotal event occurred in Galois's 1832 letters to Auguste Chevalier, where he implicitly outlined group actions on roots as permutations preserving algebraic relations, encapsulating his theory just before his death. Augustin-Louis Cauchy formalized permutation groups in the 1840s, treating them as abstract systems of substitutions. In his 1845 memoir Mémoire sur le nombre des valeurs qu'une fonction peut acquérir, lorsqu'on y permute de toutes les manières possibles les quantités qu'elle renferme, Cauchy defined permutation multiplication and explored their algebraic structure, including decompositions into cycles and disjoint cycles. He introduced conjugacy classes as sets of permutations related by simultaneous relabeling, proving that such classes partition the group and share cycle types, which became essential for classifying permutation behaviors.

Modern Developments

In the second half of the 19th century, Camille made substantial advances in the theory of permutation groups. In works culminating in his 1879 treatise Traité des substitutions et des équations algébriques, developed a systematic treatment of substitution groups, introducing concepts such as primitive and imprimitive actions, for groups, and the Jordan-Hölder theorem. His abstraction of permutation groups from concrete substitutions paved the way for modern abstract . In the late 19th and early 20th centuries, William Burnside advanced the understanding of through his work on and , which provided tools for analyzing subgroup structures and early attempts at classification. Burnside's , which states that if a Sylow p-subgroup P of a G satisfies P ⊆ Z(N_G(P)), then G has a , relied on to establish conditions for the existence of complementing Sylow subgroups. These ideas, developed in his foundational text, laid groundwork for later classifications by linking to broader group-theoretic properties. A major milestone in the mid-20th century was the O'Nan-Scott theorem, formulated in the 1970s and 1980s, which provides a of finite groups into five broad types: affine, almost , diagonal, product action, and twisted types. This theorem, originally outlined in unpublished notes by Michael O'Nan and Scott, categorizes groups based on the structure of their socles and stabilizers, reducing the study of such groups to cases involving groups or their products. The has been instrumental in applications to and , such as enumerating designs and graphs with specified symmetries. In the 1980s, Michael Aschbacher provided key refinements to the O'Nan-Scott theorem, resulting in what is now known as the Aschbacher-O'Nan-Scott theorem, which expands the classification to eight types by incorporating extensions like the holomorph of vector spaces and correcting earlier formulations, providing a more precise structural description reliant on the . Aschbacher's contributions clarified the twisted wreath product case and integrated quasiprimitive groups, enhancing the theorem's completeness for actions. Concurrently, computational methods emerged with Charles Sims' development of the Schreier-Sims in the 1970s, which computes a and strong generating set (BSGS) for a permutation group given by generators, enabling efficient determination of the group's order and membership testing. The constructs a stabilizer chain using Schreier's lemma to generate representatives, with a polynomial in the degree for groups of bounded non-triviality, making it foundational for software implementations. This breakthrough shifted permutation group theory toward algorithmic approaches, facilitating the verification of theoretical classifications on large examples. Computational group theory has flourished through systems like and , which implement Schreier-Sims variants and O'Nan-Scott reductions to construct and analyze permutation groups up to high degrees, supporting research in random and subgroup lattices.

References

  1. [1]
    Permutation Groups - Department of Mathematics at UTSA
    Nov 17, 2021 · A permutation group is a group G whose elements are permutations of a given set M and whose group operation is the composition of ...
  2. [2]
    [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. 3. Note: We'll focus specifically on the case when A = 11 ...
  3. [3]
    [PDF] 8 Groups of Permutations - UC Berkeley math
    Theorem (Cayley's Theorem). Every group is isomorphic to a group of permutations. The proof of Cayley's theorem requires a definition and a lemma. Definition.
  4. [4]
    [PDF] Lecture 2.3: Symmetric and alternating groups
    Loosely speaking, a symmetric group is the collection of all n! permutations of n objects. Alternating groups are similar. Thus, we will study permutations, and ...
  5. [5]
    [PDF] The Alternating Groups - DSpace@MIT
    The alternating group An is the group of even permutations in Sn. Our object is to prove. Theorem. If n ≥ 5, the alternating group An is a simple group.
  6. [6]
    AATA Permutation Groups - Abstract Algebra: Theory and Applications
    Permutation groups are central to the study of geometric symmetries and to Galois theory, the study of finding solutions of polynomial equations.
  7. [7]
    [PDF] Permutation Groups in NC - University of Oregon
    Abstract. We show that the basic problems of permutation group manipulation admit e cient parallel solutions. Given a permutation group G by a list of ...
  8. [8]
    [PDF] §3.6 Permutation Groups - University of South Carolina
    Any subgroup of the symmetric group Sym(S) is called a permutation group. Every group G is isomorphic to a permutation group. Proof: Given a ∈ G, define λa : G ...
  9. [9]
  10. [10]
    [PDF] The symmetric group
    Definition 1.3. The symmetric group Sn is the group Perm({1,...,n}) of all permutations on the first n integers.
  11. [11]
  12. [12]
    Basic definitions and facts
    Syllabus: In this section we give the precise definition of an abstract group and recall some basic examples. We also recall basic properties of permutations ...Missing: finiteness | Show results with:finiteness
  13. [13]
    [PDF] Permutation - OU Math
    Permutations. 1. Definition (Permutation). A permutation of a set A is a bijective function f : A → A. The set of all permutations of A is denoted by Perm(A).
  14. [14]
    [PDF] Finite sets, counting and group theory - Purdue Math
    Theorem 5.9 (Lagrange). If H ✓ G is a subgroup of a finite group, then |G| = |H|·|G/H| In particular, the order of H divides the order of G. Proof. By the ...
  15. [15]
    one-line notation for permutations - PlanetMath.org
    Mar 22, 2013 · So now we can translate the permutation into cycle notation: π=(1)(274)(56) π = ( 1 ) ⁢ ( 274 ) ⁢ ( 56 ) . If we are willing to allow words of ...
  16. [16]
    1.4 Permutations | MATH0007: Algebra for Joint Honours Students
    One way to do this is two row notation in which we write the numbers 1 to n in a row and then write σ(i) σ ( i ) beneath i: (12⋯n−1nσ(1)σ(2)⋯σ(n−1)σ(n)) ( 1 2 ⋯ ...
  17. [17]
    15.3: Permutation Groups - Mathematics LibreTexts
    Aug 16, 2021 · Cycle Notation. A second way of describing a permutation is by means of cycles, which we will introduce first ...
  18. [18]
    26.13 Permutations: Cycle Notation
    In cycle notation, the elements in each cycle are put inside parentheses, ordered so that σ ⁡ ( j ) immediately follows j or, if j is the last listed element ...
  19. [19]
    [PDF] group actions - keith conrad
    What we have been calling a group action could be called a left group action, while a right group action, denoted xg, has the properties xe = x and (xg1)g2 = x( ...
  20. [20]
    [PDF] Imprimitive Permutation Groups - UP FAMNIT
    Any other block is nontrivial. If G has a nontrivial block then it is imprimitive. If G is not imprimitive, we say that G is primitive. Note that if B ...
  21. [21]
    Permutation -- from Wolfram MathWorld
    ### Summary of Permutation Composition from MathWorld
  22. [22]
    [PDF] Introduction to Abstract Algebra (Math 113)
    3.4 Permutation Groups and Group Actions . ... We have an analogous concept in group theory. Definition. Let (G, ∗) be a group. A subgroup of G is a ...
  23. [23]
    Permutation Group -- from Wolfram MathWorld
    A permutation group is a finite group G whose elements are permutations of a given set and whose group operation is composition of permutations in G.
  24. [24]
    [PDF] Group Theory - James Milne
    There are over a hundred exercises, many with solutions. BibTeX information. @misc{milneGT, author={Milne, James S.}, title={Group Theory ...
  25. [25]
    SymmetricGroup
    The symmetric group on n letters, written S n , has as its elements the n! distinct bijections from an n-element set (often taken to be {1...n}) to itself.
  26. [26]
    [PDF] 13. Symmetric groups
    The symmetric group Sn is the group of bijections of {1,...,n} to itself, also called permutations of n. things. A standard notation for the permutation that ...Missing: S_n | Show results with:S_n
  27. [27]
    [PDF] Math 412. The Symmetric Group Sn.
    The Symmetric Group Sn. DEFINITION: The symmetric group Sn is the group of bijections from any set of n objects, which we usu- ally call simply {1,2,...,n} ...
  28. [28]
    [PDF] Sign of permutations - Keith Conrad
    Our starting point is that all permutations in Sn are products of transpositions. Theorem 1.1. For n ≥ 2, Sn is generated by its transpositions. Proof. Here are ...Missing: S_n | Show results with:S_n
  29. [29]
    [PDF] The Alternating Group
    An is called the alternating group. An important feature of the alternating group is that, unless n = 4, it is a simple group. A group G is said to be simple ...
  30. [30]
    [PDF] Chapter 1: Abstract Group Theory - Rutgers Physics
    Definition: The alternating group An ⊂ Sn is the subgroup of Sn of even permuta- ... describe below, An for n ≥ 5 are all simple groups. Moreover, |An| is ...
  31. [31]
    [PDF] Lecture 15 - Math 5111 (Algebra 1)
    Nov 2, 2020 · Theorem (Alternating Group) The alternating group An is the kernel of the sign map and is therefore is a normal subgroup of Sn. Explicitly, An ...Missing: A_n S_n simple
  32. [32]
    [PDF] SIMPLICITY OF An - KEITH CONRAD
    A finite group is simple if it's nontrivial and its only normal subgroups are the trivial subgroup and the whole group. For n ≥ 5, the group An is simple.
  33. [33]
    [PDF] Math 412. Symmetric Group: Answers.
    The order of the four permutations that are products of disjoint transpositions is 2. (8) An example of a cyclic subgroup of order 2 is h(1 2)i = {e, (1 2)}.
  34. [34]
    Cyclic Group -- from Wolfram MathWorld
    ### Summary of Cyclic Group \( C_n \) Representation and Properties
  35. [35]
    3.3: Dihedral Groups (Group of Symmetries)
    ### Summary of Dihedral Groups as Permutation Groups
  36. [36]
    Dihedral Group -- from Wolfram MathWorld
    The dihedral group D_n is the symmetry group of an n-sided regular polygon for n>1. The group order of D_n is 2n. Dihedral groups D_n are non-Abelian ...
  37. [37]
    group actions and homomorphisms - PlanetMath.org
    Mar 22, 2013 · A group action of G on X is a mapping where g⋅x is defined. A group homomorphism F:G→SX maps g to fg. A group action can be induced by a  ...
  38. [38]
    [PDF] Group Actions and Permutations Representations
    permutation representation associated to the group action of. G on A ... 3 An action is said to be faithful if its kernel is the identity. Kevin James.
  39. [39]
    [PDF] Abstract Algebra I
    Since the kernel of an action is the same as the kernel of the associated permutation representation, it is a normal subgroup of G. Two group elements induce ...
  40. [40]
    [PDF] GROUPS ACTING ON A SET 1. Left group actions Definition 1.1 ...
    We now consider the same example in a slightly more explicit framework. Example 1.3. Suppose that G = S4, the group of permutations on the set S = {1,2,3,4}.
  41. [41]
    [PDF] More about Permutations and Symmetry Groups - Purdue Math
    Theorem 10.7 (Orbit-Stabilizer theorem I). Given a transitive subgroup G ✓. Sn, let H be the stabilizer of some element i, then |G| = n|H|. Let us re ...
  42. [42]
    [PDF] The Orbit Stabilizer Theorem
    Dec 9, 2002 · The Orbit Stabilizer Theorem: Let G be a finite group of permutations of a set S. Then for any i from S, |G| = |orbG(i)|·|stabG(i)|. Example:
  43. [43]
    [PDF] finding minimal permutation representations of finite groups
    Apr 21, 2008 · Any permutation representation is a disjoint union of orbits. Definition 2.1. The core K(H) of a subgroup H of G is the maximal normal subgroup ...<|control11|><|separator|>
  44. [44]
    Permutation Groups | SpringerLink
    Free delivery 14-day returnsThis text can serve as an introduction to permutation groups in a course at the graduate or advanced undergraduate level, or for self- study.
  45. [45]
    [PDF] Transitive group actions - Keith Conrad
    Definition 4.1. An action of a group G on a set X, with |X| ≥ 2, is called doubly transitive when, for all ordered pairs of distinct elements (x, x0) and (y, y ...
  46. [46]
    A classification of primitive permutation groups with finite stabilizers
    Jun 15, 2015 · A transitive group G is primitive if and only if for all α ∈ Ω the point stabilizer G α is a maximal subgroup of G. Note that if G is transitive ...
  47. [47]
    [PDF] Primitive permutation groups 1 The basics 2 Minimal normal ...
    Aug 27, 2004 · The action of G on the set of right cosets of H by right multiplication is primitive if and only if H is a maximal proper subgroup of G. Thus, ...
  48. [48]
    [PDF] Extremely primitive sporadic and alternating groups - SEIS
    Nov 4, 2011 · Abstract. A non-regular primitive permutation group is said to be extremely primitive if a point stabilizer acts primitively on each of its ...
  49. [49]
    [PDF] Imprimitive permutations in primitive groups
    The permutation g is contained in an imprimitive permutation group with k blocks of size m if and only if the cycle partition of g on {1,...,km} (with k,m > ...
  50. [50]
    On the O'Nan-Scott theorem for finite primitive permutation groups
    We give a self-contained proof of the O'Nan-Scott Theorem for finite primitive permutation groups.
  51. [51]
    [PDF] 21 Symmetric and alternating groups
    1) Every finite abelian group is solvable. 2) For n ≥ 5 the symmetric group Sn has a composition series. {(1)} ⊆ An ⊆ Sn and so Sn is not solvable. 23.3 ...Missing: S_n | Show results with:S_n
  52. [52]
    [PDF] Basic Theory of Affine Group Schemes - James Milne
    Mar 11, 2012 · A group G is said to be solvable if the derived series. G DG D. 2. G terminates with 1. For example, if n 5, then Sn (symmetric group on n ...
  53. [53]
    [PDF] Constructing Transitive Permutation Groups
    This question is formulated in the language of invariants – at this time there was no formal definition of a permutation group – and what it asks for are ...
  54. [54]
    [PDF] arXiv:2305.08405v1 [math.GR] 15 May 2023
    May 15, 2023 · The next class of groups we consider are permutation groups, all of whose non- ... has a composition series all of whose factors are cyclic groups ...
  55. [55]
    [PDF] Frobenius groups - Huskie Commons
    A finite group G is called a Frobenius group if there is a non-trivial subgroup H of G, called the Frobenius complement, such that for all g ∈ G\H we have gHg−1 ...Missing: semiregular | Show results with:semiregular
  56. [56]
    [PDF] Chapter 7. Frobenius Groups - RPTU
    (a) To use standard group theory terminology, the theorem says that the Frobenius kernel is a normal complement of H in G and that G is an internal semi-direct ...
  57. [57]
    (PDF) The Schreier-Sims algorithm - ResearchGate
    This representation helps us to calculate the grouporder, list the group elements, generate random elements, test for group mem-bership and store group elements ...
  58. [58]
    [PDF] arXiv:1812.03346v2 [math.RA] 15 Dec 2018
    Dec 15, 2018 · Our technique is an analogue to the Schreier-Sims algorithm for permutation groups and is a by-product of Frobenius reciprocity. 1. Introduction.Missing: recognition | Show results with:recognition
  59. [59]
    [PDF] A polynomial-time theory of black-box groups I 1 Introduction
    Charles Sims pioneered the design of efficient algorithms for the rudimentary tasks of managing permutation groups (membership, order, normal closure) based on ...
  60. [60]
    [PDF] Oligomorphic permutation groups - Queen Mary University of London
    Oligomorphic groups satisfy a rather different kind of finiteness condi- tion; paradoxically, one which makes them “large” rather than “small”. A permutation ...
  61. [61]
  62. [62]
    [PDF] Valued Constraint Satisfaction in Structures with an Oligomorphic ...
    Page 1. Valued Constraint Satisfaction in. Structures with an Oligomorphic. Automorphism Group ... theory ... The CSP of B (denoted. CSP(B)) is the problem of ...
  63. [63]
    Finite Permutation Groups - ScienceDirect.com
    Finite Permutation Groups. Book • 1964. Author: HELMUT WIELANDT. Finite ... PDF version. Ways of reading. No information about appearance modifiability is ...
  64. [64]
    [PDF] The Evolution of Group Theory: A Brief Survey - Israel Kleiner
    Mar 14, 2004 · In 1854 Cayley, in a paper entitled "On the theory of groups, as depending on the symbolic equation @" = 1," gave the first abstract ...
  65. [65]
    [PDF] Early group theory in the works of Lagrange, Cauchy, and Cayley
    Aug 19, 2010 · ... Lagrange (1736–1813), the first to suggest a relation between permutations and the solution of equations by radicals. In Section 2, we ...
  66. [66]
    Paolo Ruffini (1765 - 1822) - Biography - MacTutor
    Quick Info. Paolo Ruffini was an Italian mathematician who gave a proof that the quintic equation could not be solved earlier than Abel.
  67. [67]
    [PDF] Abel Answers the Question of the Quintic - IMA
    Feb 20, 2025 · In his proof, Ruffini employed an important idea from La- grange's 1770 paper, namely that the algebraic solvability of equations was strongly ...
  68. [68]
    [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 Mémoire ...
  69. [69]
    [PDF] galois-testamentary.pdf
    Letter to Auguste Chevalier. My dear friend,. Paris, 29 May 1832. I have done several new things in analysis. Some concern the theory of equations, others ...
  70. [70]
    [PDF] Early Group Theory in the Work of A.-L. Cauchy
    Jul 1, 2025 · In Section 1, we examine excerpts from his 1845 work which explain the notion of permutation multiplication and its basic properties. With these ...
  71. [71]
    The mathematical life of Cauchy's group theorem - ScienceDirect.com
    Cauchy included a variety of results about groups of permutations—conjugate systems of substitutions—in his great 1845 memoir, but he begins the section ...Missing: formalization | Show results with:formalization
  72. [72]
    [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.
  73. [73]
    The O'Nan–Scott Theorem (Chapter 4) - Permutation Groups
    The socle of a finite group G is the product of the minimal normal subgroups of G. (The original meaning of the word is 'the base on which a statue stands'.) ...
  74. [74]
    On an Algorithm for Finding a Base and a Strong Generating Set for ...
    Abstract. This paper deals with the problem of finding a base and strong gener- ating set for the group generated by a given set of permutations.
  75. [75]
    GAP (ref) - Chapter 43: Permutation Groups
    43.5-2 SocleTypePrimitiveGroup. ‣ SocleTypePrimitiveGroup ( G ), ( attribute ). returns the socle type of the primitive permutation ...
  76. [76]
    Permutation Group - Magma Computational Algebra System
    A software package designed to solve computationally hard problems in algebra, number theory, geometry and combinatorics.Missing: solvable derived affine