Fact-checked by Grok 2 weeks ago

Group action

In , a group action is a fundamental concept in where a group G operates on a set X through transformations that preserve the group's , effectively describing symmetries or permutations induced by the group elements. Formally, a group action is a \phi: G \times X \to X, often denoted (g, x) \mapsto g \cdot x, satisfying two axioms: the e \in G acts trivially, so e \cdot x = x for all x \in X, and the action is compatible with group multiplication, so g \cdot (h \cdot x) = (gh) \cdot x for all g, h \in G and x \in X. This turns X into a G-set, allowing the group to "act" on it in a consistent manner. Group actions generalize the notion of symmetry groups, such as the acting on the vertices of a by rotations and reflections, and are essential for analyzing how groups interact with geometric, algebraic, or combinatorial objects. Central to the theory are orbits and stabilizers: the orbit of x \in X is the set \{ g \cdot x \mid g \in G \}, representing all points reachable from x via the action, while the stabilizer of x is the \{ g \in G \mid g \cdot x = x \}, consisting of that fix x. The orbit-stabilizer establishes a key relationship: if G is finite, the size of the orbit equals the index of the stabilizer in G, providing a tool to compute cardinalities and classify actions as transitive (single orbit), free (trivial stabilizers), or faithful (injective action map). Beyond basic properties, group actions enable profound applications across mathematics. In , the acts on the roots of a , with orbits corresponding to irreducible factors, linking algebra to field extensions. In , uses fixed points of group elements to count distinct objects under symmetry, such as necklace colorings or molecular configurations. Actions also extend to topological and geometric settings, where continuous group actions model symmetries in manifolds or spaces, influencing , , and physics via symmetry groups in and . These concepts, first systematically explored in the context of permutation groups and Galois correspondences in the , remain indispensable for proving theorems like Sylow's in finite group theory and understanding modular forms.

Definitions

Left actions

A left action of a group G on a set X is a map G \times X \to X, denoted (g, x) \mapsto g \cdot x, satisfying the identity condition e \cdot x = x for all x \in X, where e is the of G, and the compatibility condition (gh) \cdot x = g \cdot (h \cdot x) for all g, h \in G and x \in X. These axioms ensure that the action respects the group structure, allowing elements of G to "transform" points in X in a consistent manner that mirrors group . The notation for such an action is commonly expressed as G acting on X, with g \cdot x often simplified to gx. This shorthand facilitates concise descriptions in algebraic contexts, emphasizing the operational nature of the action. The structured pair (X, \cdot), where the action is specified, is termed a G-set, encapsulating both the underlying set and the group operation on it. The concept of group actions has roots in 19th-century permutation group theory, with the modern abstract definition formalized in early 20th-century developments in group theory.

Right actions

A right of a group G on a set X is a X \times G \to X, written (x, g) \mapsto x \cdot g, that satisfies two axioms: the e \in G acts trivially, so x \cdot e = x for all x \in X, and the respects the group operation from the right, so (x \cdot g) \cdot h = x \cdot (gh) for all x \in X and g, h \in G. This contrasts with the standard left , where the group operation is applied from the left: g \cdot (h \cdot x) = (gh) \cdot x. The notation for right actions often uses x^g to denote x \cdot g, or simply xg, emphasizing the rightward application. Right actions arise naturally in contexts like permutation representations where the order of group elements matters for composition, such as when viewing as acting on the right in certain studies. However, right actions are equivalent to left actions up to inversion of group elements. Specifically, given a right x \cdot g, one defines a corresponding left by g \cdot x = x \cdot g^{-1}; this map satisfies the left axioms because inversion turns the right compatibility into left compatibility: g \cdot (h \cdot x) = (x \cdot h^{-1}) \cdot g^{-1} = x \cdot (h^{-1} g^{-1}) = x \cdot (gh)^{-1} = (gh) \cdot x. The converse holds similarly, establishing a between right G-actions on X and left G-actions on X. When the group G is abelian, left and right actions coincide in a stronger sense because the inversion map g \mapsto g^{-1} is itself a group homomorphism: (gh)^{-1} = g^{-1} h^{-1} = h^{-1} g^{-1}. This makes the equivalence canonical and preserves the group structure directly.

Fundamental Properties

A group action of a group G on a set X is called transitive if for every pair of elements x, y \in X, there exists an element g \in G such that g \cdot x = y. Equivalently, the action is transitive if and only if X forms a single orbit under the action of G. A stronger notion is that of a doubly transitive action, where for any two ordered pairs of distinct elements (x_1, x_2), (y_1, y_2) \in X \times X with x_1 \neq x_2 and y_1 \neq y_2, there exists g \in G such that g \cdot x_1 = y_1 and g \cdot x_2 = y_2. Every doubly transitive action is transitive, but the converse does not hold in general. An action is termed faithful if the associated homomorphism \phi: G \to \mathrm{Sym}(X) from G to the symmetric group on X is injective, meaning the kernel is trivial and only the identity element of G fixes every point in X. In other words, if g \cdot x = x for all x \in X implies g is the identity. Finally, an action is regular if it is both faithful and transitive, and in such cases, the cardinalities satisfy |G| = |X|. Equivalently, for any x, y \in X, there is exactly one g \in G such that g \cdot x = y. Regular actions provide a canonical way to realize G as a permutation group on a set of the same size.

Primitivity

In group theory, a subset B \subseteq X of the set X on which a group G acts is called a block if $1 < |B| < |X| and for every g \in G, either gB = B or gB \cap B = \emptyset. Such blocks correspond to the parts of nontrivial partitions of X that are preserved by the action of G. A group action of G on a set X is primitive if it is transitive and admits no nontrivial blocks. Equivalently, the only G-invariant partitions of X are the trivial ones consisting of singletons or the full set \{X\}. Primitivity thus captures a notion of "maximal transitivity" in the sense that the action cannot be refined into a coarser transitive action on a system of blocks without losing transitivity on X. A key characterization of primitive actions is that a transitive action of G on X is primitive if and only if the stabilizer G_x of any point x \in X is a maximal subgroup of G. This equivalence highlights the structural rigidity of primitive actions: the point stabilizers leave no room for intermediate subgroups that could induce nontrivial block systems. In the context of permutation groups, this means that primitive subgroups of the S_n arise precisely as the images of transitive actions where stabilizers are maximal. Primitive actions are particularly significant for simple groups, as their faithful primitive representations embed the group as a primitive permutation subgroup of some symmetric group S_n, providing insights into the embedding problem and the classification of finite simple groups via permutation representations.

Orbits and Stabilizers

Orbits and invariant subsets

The orbit of an element x \in X under a group action of G on X is the set \Orb_G(x) = \{ g \cdot x \mid g \in G \}, consisting of all elements of X that can be obtained by acting on x with elements of G. This set is the equivalence class of x under the relation \sim defined by x \sim y if and only if there exists g \in G such that y = g \cdot x; the relation \sim is an equivalence relation on X. The induced by the group action partitions the set X into disjoint , decomposing X as a of these . Each is stable under the action, and the restriction of the action to an yields a transitive action on that . A Y \subseteq X is under the group action if g \cdot Y = Y for all g \in G. Equivalently, every is a union of , and conversely, every union of is . Thus, the collection of corresponds precisely to the of X that are saturated with respect to the partition into .

Stabilizers and fixed points

In group theory, given a group G acting on a set X, the stabilizer of an x \in X, denoted \operatorname{Stab}_G(x) or simply \operatorname{Stab}(x), is the consisting of all of G that fix x under the action. Specifically, \operatorname{Stab}_G(x) = \{ g \in G \mid g \cdot x = x \}. This set forms a of G because it is nonempty (containing the ), closed under the group operation (if g \cdot x = x and h \cdot x = x, then (gh) \cdot x = g \cdot (h \cdot x) = g \cdot x = x), and closed under inverses (if g \cdot x = x, then g^{-1} \cdot x = x by applying g^{-1} to both sides of the equation). Different of X may have distinct , and the stabilizer captures the "symmetry" preserving a particular point. For a fixed group element g \in G, the fixed points of g are the elements of X that remain unchanged when acted upon by g. Formally, the fixed-point set of g, denoted \operatorname{Fix}_G(g) or X^g, is defined as \operatorname{Fix}_G(g) = \{ x \in X \mid g \cdot x = x \}. This set measures the extent to which g acts trivially on X, and it always includes the points stabilized by g alone. The e \in G has \operatorname{Fix}_G(e) = X, as it fixes every point by definition. An action of G on X is termed fixed-point-free if the only element of G with nonempty fixed-point set beyond the trivial case is the identity, meaning \operatorname{Fix}_G(g) = \emptyset for all g \in G \setminus \{e\}. Equivalently, such an action is called free, where every non-identity element moves all points in X, implying that stabilizers are trivial (\operatorname{Stab}_G(x) = \{e\} for all x \in X). Free actions are fundamental in studying coverings and quotients, as they ensure the action is "as injective as possible" on points. When the set X is itself a group (say H) and G acts on H by conjugation, the stabilizer of an element h \in H coincides with the centralizer of h in G, defined as \{ g \in G \mid g h g^{-1} = h \}, or equivalently \{ g \in G \mid g h = h g \}. This specializes the general concept to actions preserving the group structure, highlighting commutativity relations within G.

Orbit-stabilizer theorem

The - theorem establishes a fundamental relationship between the of an under a group and the of that . For a group G acting on a set X, and for any x \in X, the cardinality of the \operatorname{Orb}_G(x) equals the index of the \operatorname{Stab}_G(x) in G: |\operatorname{Orb}_G(x)| = [G : \operatorname{Stab}_G(x)]. If G is finite, this simplifies to |\operatorname{Orb}_G(x)| = |G| / |\operatorname{Stab}_G(x)|. The proof proceeds by constructing a bijection between the set of left cosets G / \operatorname{Stab}_G(x) and \operatorname{Orb}_G(x). Define the map \phi: G / \operatorname{Stab}_G(x) \to \operatorname{Orb}_G(x) by \phi(g \operatorname{Stab}_G(x)) = g \cdot x. This is well-defined because if g \operatorname{Stab}_G(x) = g' \operatorname{Stab}_G(x), then g'^{-1} g \in \operatorname{Stab}_G(x), so g' \cdot x = g \cdot x. It is surjective since every element in the orbit is g \cdot x for some g \in G. It is injective because if \phi(g \operatorname{Stab}_G(x)) = \phi(g' \operatorname{Stab}_G(x)), then g \cdot x = g' \cdot x, implying g^{-1} g' \in \operatorname{Stab}_G(x) and thus g \operatorname{Stab}_G(x) = g' \operatorname{Stab}_G(x). For a right action, the proof uses right cosets analogously. When G is finite and the action is transitive (i.e., X is a single ), the implies that |X| divides |G|, since |\operatorname{Stab}_G(x)| divides |G| by . In a regular action, where the action is both transitive and (stabilizers are trivial), it follows that |X| = |G|. For infinite groups, the ensures that the of the equals the of the [G : \operatorname{Stab}_G(x)], providing a cardinal arithmetic analogue.

Burnside's lemma

Burnside's lemma, also known as the Cauchy–Frobenius lemma, is a fundamental result in group theory that enumerates the number of orbits in a on a set. For a G acting on a X, the lemma states that the number of orbits |X / G| is equal to the average number of points fixed by each group element: |X / G| = \frac{1}{|G|} \sum_{g \in G} |\mathrm{Fix}(g)|, where \mathrm{Fix}(g) = \{ x \in X \mid g \cdot x = x \} denotes the fixed points of g. This formula aggregates the fixed-point data over the entire group to yield a global count of distinct orbits under the action. The proof proceeds by double counting the set of pairs (g, x) \in G \times X such that g \cdot x = x. The size of this set is \sum_{g \in G} |\mathrm{Fix}(g)|. Alternatively, fixing x \in X first, the number of such g is |\mathrm{Stab}(x)|, so the total is \sum_{x \in X} |\mathrm{Stab}(x)|. By the orbit-stabilizer theorem, |\mathrm{Stab}(x)| = |G| / |\mathrm{Orbit}(x)|, yielding \sum_{x \in X} |G| / |\mathrm{Orbit}(x)| = |G| \sum_{x \in X} 1 / |\mathrm{Orbit}(x)|. The inner sum equals the number of orbits, since each orbit contributes exactly once regardless of its size. Thus, |X / G| = \frac{1}{|G|} \sum_{g \in G} |\mathrm{Fix}(g)|. For infinite groups, a version of holds when G is a compact equipped with a normalized \mu (satisfying \mu(G) = 1). If G acts continuously on a discrete set X with finitely many orbits, the number of orbits is \int_G |\mathrm{Fix}(g)| \, d\mu(g). This generalizes the finite average, leveraging the invariance of the under group translations. A significant combinatorial application of is Pólya's enumeration theorem, which originated as a method to count the number of distinct colorings of a structure up to group symmetries by incorporating the cycle structures of group elements into a known as the .

Examples and Applications

Classical examples

One of the most fundamental examples of a is the natural action of the S_n on the X = \{1, 2, \dots, n\}, where each \sigma \in S_n acts by \sigma \cdot i = \sigma(i) for i \in X. This action is faithful, meaning the of the corresponding S_n \to \mathrm{Sym}(X) is trivial, as distinct permutations move elements differently. It is also transitive, since any two elements in X can be mapped to each other by some permutation, and primitive for n \geq 2, as the only blocks are trivial singletons or the full set. Another classical example involves the C_n = \langle r \rangle of order n, acting regularly on the set of n-th of unity in the complex numbers, denoted \mu_n = \{ e^{2\pi i k / n} \mid k = 0, 1, \dots, n-1 \}, via : r^k \cdot \zeta = r^k \zeta for \zeta \in \mu_n. Here, \mu_n itself forms a under , isomorphic to C_n, and the action is regular because the of any is trivial and the of any is the entire set. The D_n of order $2n, consisting of the symmetries of a regular n- (rotations and ), acts on the set of n vertices of the by rigidly mapping the figure to itself. This action is transitive, as any vertex can be mapped to any other via a suitable or . Furthermore, it is if and only if n is prime, since non-prime n allows nontrivial blocks corresponding to divisors of n. A simple yet illustrative example is the trivial action of any group G on an arbitrary nonempty set X, defined by g \cdot x = x for all g \in G and x \in X. In this case, every element of X is a fixed point, so each singleton \{x\} is both an orbit and a fixed set, and the action has trivial stabilizers everywhere. Finally, every group G acts on itself by conjugation, where g \cdot h = g h g^{-1} for g, h \in G. The orbits of this action are precisely the conjugacy classes of G, which partition G into subsets of elements sharing the same cycle type (in the symmetric group case) or equivalent under inner automorphisms more generally. The stabilizer of h is the centralizer C_G(h), and this action is useful for studying the structure of G via its class equation.

Combinatorial applications

Group actions find prominent applications in , particularly in enumerating distinct objects up to through the use of , which counts the orbits of the action. This approach is essential for problems where symmetries, such as rotations or reflections, identify equivalent configurations, allowing the computation of inequivalent structures by averaging the fixed points over the group elements. A classic example is the counting of necklaces, where the cyclic group C_n acts on the set of colorings of n beads with a fixed number of colors. Here, two colorings are considered the same if one can be obtained from the other by , and determines the number of distinct necklaces by summing the fixed colorings for each group element and dividing by the group order. For instance, with c colors, the formula yields the number of rotationally distinct arrangements, providing a practical tool for such symmetric enumerations. In , the \operatorname{Aut}(G) of a G acts on the set of proper k-colorings, where the orbits correspond to colorings that are inequivalent under the graph's symmetries. Applying counts these orbits, yielding the number of distinct colorings up to , which is useful for classifying structures. Burnside's lemma also appears in the of orbits within combinatorial , where group actions on labeled structures help count unlabeled by considering symmetries in generating functions. Historically, Burnside introduced the in his 1897 book Theory of Groups of Finite Order, applying it to combinatorial problems in theory.

Morphisms of Actions

Homomorphisms between G-sets

A homomorphism between two G-sets X and Y, also known as a G-equivariant map, is a \phi: X \to Y satisfying \phi(g \cdot x) = g \cdot \phi(x) for all g \in G and x \in X. This condition ensures that \phi intertwines the actions of G on X and Y, preserving the group action structure. The collection of all G-sets, together with these G-s as morphisms, forms a denoted \mathbf{Set}^G. Any group action of G on a set X induces a representation, which is a \rho: G \to \mathrm{Sym}(X) defined by \rho(g)(x) = g \cdot x for all g \in G and x \in X, where \mathrm{Sym}(X) is the on X. This homomorphism embeds the action into the of permutation groups, allowing the study of actions via their images in \mathrm{Sym}(X). The kernel of a group action on X is the set \ker(\rho) = \{ g \in G \mid g \cdot x = x \ \forall x \in X \}, which coincides with the kernel of the induced homomorphism \rho: G \to \mathrm{Sym}(X). As the kernel of a group homomorphism, \ker(\rho) is a normal subgroup of G. An action is faithful if and only if this kernel is trivial. By the first isomorphism theorem for groups, the image \mathrm{im}(\rho) is a of \mathrm{Sym}(X) isomorphic to the G / \ker(\rho). This isomorphism identifies the of G modulo its with a acting on X. G-equivariant maps preserve , mapping the orbit of x \in X into the orbit of \phi(x) \in Y.

Isomorphisms and conjugacy

A G-isomorphism between two G-sets (X, \cdot) and (Y, \circ) is a bijective map f: X \to Y that is a G-homomorphism, meaning f(g \cdot x) = g \circ f(x) for all g \in [G](/page/G) and x \in X, with the additional property that its inverse f^{-1}: Y \to X is also a G-homomorphism. Since f is bijective, the equivariance of the inverse follows automatically from the group action axioms, ensuring that isomorphisms preserve the entire structure of the actions. Two group actions on sets X and Y are said to be if their corresponding G-sets are isomorphic, i.e., there exists a G-isomorphism between them. This classifies actions up to structural similarity, preserving key features such as the orbit-stabilizer theorem: isomorphic G-sets have corresponding orbits of equal and stabilizers that are isomorphic as subgroups of G. For instance, the left regular action of G on itself is equivalent to any transitive free action on a set of |G|. Conjugate actions provide another way to relate actions of the same group G on a fixed set X. Given an \cdot : G \times X \to X and an \alpha \in \Aut(G), the conjugate \cdot_\alpha is defined by g \cdot_\alpha x = \alpha(g) \cdot x for all g \in G and x \in X. This defines a new group , as \alpha preserves the group operation, and it twists the original by relabeling elements of G via \alpha. The conjugation of G on itself, defined by g \cdot x = g x g^{-1} for g, x \in G, is a related example, yielding orbits as conjugacy classes. Cayley's theorem illustrates these concepts through the regular action: every group G embeds as a of the \Sym(G) via the left regular action \lambda: G \to \Sym(G) defined by \lambda(g)(h) = gh for g, h \in G, yielding a faithful transitive action isomorphic to the standard permutation representation. This embedding is unique up to conjugacy in \Sym(G), meaning any two such regular embeddings differ by conjugation by some permutation in \Sym(G), reflecting the equivalence of regular actions under relabeling of the set.

Advanced Variants

Topological group actions

A G acts continuously on a X if the action map \mu: G \times X \to X, defined by \mu(g, x) = g \cdot x, is continuous with respect to the on G \times X. This joint continuity ensures that the action respects the topological structures of both G and X, distinguishing it from merely set-theoretic actions by requiring that nearby group elements act in a controlled manner on points in X. In many cases, such as when G is a and X is locally compact, this implies that each fixed g \in G acts via a on X, though the converse requires additional assumptions like separate continuity. Topological transitivity extends the algebraic notion of to the continuous setting: an is topologically transitive if, for every pair of nonempty open sets U, V \subseteq X, there exists g \in G such that g \cdot U \cap V \neq \emptyset, or equivalently, the G \cdot U is dense in X. In compact spaces, this is equivalent to the existence of at least one dense . Topological transitivity captures mixing or ergodic-like behavior in dynamical systems arising from group s, where orbits densely fill the space without collapsing to finite sets as in the discrete case. A continuous action is proper if the map G \times X \to X \times X, (g, x) \mapsto (x, g \cdot x), is a proper map, meaning that the preimage of every compact subset of X \times X is compact in G \times X. Equivalently, for every compact K \subseteq X, the set \{(g, x) \in G \times X \mid x \in K, g \cdot x \in K\} is compact. Proper actions ensure well-behaved spaces and stabilizers, often compact, which is crucial for constructions like orbifolds; compact groups always act properly on Hausdorff spaces. A classic example is the continuous action of the special orthogonal group SO(3), the group of 3D rotations, on \mathbb{R}^3 \setminus \{0\} by matrix-vector multiplication: for A \in SO(3) and v \in \mathbb{R}^3 \setminus \{0\}, A \cdot v = Av. This action is continuous since matrix multiplication is a polynomial map, hence continuous, and proper because SO(3) is compact. The orbits are spheres centered at the origin, illustrating how properness prevents "accumulation at infinity." In modern dynamical systems, such topological group actions underpin the study of qualitative behaviors like chaos and recurrence on non-compact spaces, extending classical ergodic theory to broader group structures.

Linear representations

A linear group action on a vector space arises when a group G acts on a V over a k in a manner that respects the linear structure of V. Specifically, the action satisfies g \cdot (a v + b w) = a (g \cdot v) + b (g \cdot w) for all g \in G, scalars a, b \in k, and vectors v, w \in V. Such actions are equivalently described by group homomorphisms \rho: G \to \GL(V), where \GL(V) is the general linear group of invertible linear endomorphisms of V; here, g \cdot v = \rho(g) v. These homomorphisms are termed linear representations of G on V. Associated to any such representation is its \chi: G \to k, defined by \chi(g) = \trace(\rho(g)), the of the linear map \rho(g). An under a linear \rho is a W \subseteq V such that \rho(g) w \in W for all g \in G and w \in W. A is irreducible if the only are \{0\} and V itself, meaning no nontrivial proper exist. Irreducible representations form the building blocks of more general representations, as they cannot be decomposed further into simpler linear actions. For finite groups G, Maschke's theorem provides a key decomposition result: if the characteristic of k does not divide the order |G|, then every finite-dimensional representation of G over k is completely reducible, meaning it decomposes as a of irreducible representations. This holds in particular over fields of characteristic zero, such as the complex numbers, but also over finite fields when the characteristic avoids dividing |G|; in cases where the characteristic divides |G|, representations may fail to be completely reducible, as seen in over fields like \mathbb{F}_p for p-groups.

Actions on groupoids

A is defined as a small in which every is an . This structure generalizes groups, which are groupoids with a single object, by allowing multiple objects connected by invertible arrows representing symmetries or equivalences between them. An action of a group G on a \mathcal{C} consists of actions of G on both the set of objects \mathrm{Ob}(\mathcal{C}) and the set of morphisms \mathrm{Mor}(\mathcal{C}), compatible with the structure of \mathcal{C}. Specifically, for each g \in G, the map on morphisms g \cdot (-) : \mathrm{Mor}(\mathcal{C}) \to \mathrm{Mor}(\mathcal{C}) satisfies s(g \cdot f) = g \cdot s(f) and t(g \cdot f) = g \cdot t(f) for source s and target t, preserves composition via g \cdot (f \circ h) = (g \cdot f) \circ (g \cdot h), and maps identities to identities g \cdot \mathrm{id}_x = \mathrm{id}_{g \cdot x}. This compatibility ensures that each g induces an of the \mathcal{C}, and the map G \to \mathrm{Aut}(\mathcal{C}) is a . Ordinary group actions on sets recover this framework when \mathcal{C} is the discrete on the set, with only identity morphisms. In geometric contexts, actions on groupoids extend to pseudogroup actions, where a pseudogroup—a collection of local homeomorphisms satisfying group-like axioms—acts on a space by generating local transformations, unifying global group actions with infinitesimal symmetries like actions. Such structures are essential for studying foliations and orbifolds, where the pseudogroup encodes local equivalence relations. Modern developments in , particularly Baez's work on groupoidification in the 2000s, reinterpret group actions on groupoids categorically to "categorify" linear algebra: vector spaces become groupoids, and linear maps become spans of groupoids, yielding insights into representation theory and quantum protocols via weak quotients and orbit-stabilizer relations. Morphisms of such actions, generalizing G-set homomorphisms, are functors intertwining the group actions.

Generalizations

Partial actions

In group theory, a partial action of a G on a set X generalizes the classical of a group by allowing the action of certain group elements to be undefined on parts of the set. Specifically, it consists of a family of subsets \{D_g \subseteq X \mid g \in G\}, called domains, together with bijections \theta_g : D_{g^{-1}} \to D_g for each g \in G, satisfying the following conditions: D_e = X and \theta_e is the map on X, where e is the group ; for all g, h \in G, the set equality D_{gh} = \theta_g(D_{h^{-1}}) holds, and the maps as \theta_{gh} = \theta_g \circ \theta_h whenever the composition is defined, i.e., on D_{h^{-1} g^{-1}}. The partial action is then denoted by g \cdot x = \theta_g(x) for x \in D_{g^{-1}}, with the action undefined otherwise. This structure ensures partial compatibility with the group , preserving bijectivity where defined. A set X equipped with such a partial is termed a partial G-set. The domains D_g often exhibit additional structure, such as forming a graded where intersections and images align with the group , enabling the partial G-set to behave like a subsystem of a full . For instance, subsets of X under the partial can naturally serve as domains in extensions or restrictions of the structure. Partial G-sets generalize full G-sets, where D_g = X for all g, and every partial embeds into a full on a larger set via constructions like the universal enveloping , which adjoins elements to make the action total while preserving the original partial dynamics. Partial actions find significant applications in , where they extend to actions on algebras, allowing groups to act partially on non-unital or rings. In this context, a partial \alpha of G on a ring R assigns to each g \in G an A_g \subseteq R (playing the role of D_g) and a ring \alpha_g : A_{g^{-1}} \to A_g, with analogous compatibility conditions. This leads to the partial skew R \rtimes^\alpha G, a generalization of the classical skew group ring that captures crossed product-like structures even when full actions are unavailable. The emerged in the late , with foundational work establishing conditions for —embedding partial actions into full actions on larger rings—and associativity of the partial skew group ring, particularly for s-unital rings. These algebraic partial actions have facilitated developments in noncommutative , where they describe extensions of rings under group symmetries that are not globally defined, and in the of simple Artinian rings via partial Galois correspondences. For rings, every partial action admits a weak enveloping action, ensuring the partial skew group ring is well-behaved and Morita equivalent to a full crossed product under certain regularity conditions. This framework, built on contributions from the 2000s onward, underscores partial actions' role in bridging set-theoretic dynamics with algebraic invariants.

Mackey functors

In equivariant homotopy theory and , Mackey functors provide a generalization of group actions by axiomatizing the , restriction, and conjugation operations that arise in the study of G-sets and representations for a G over a R. A Mackey functor M assigns to each H ≤ G an M(H), together with natural maps: restriction res^K_H : M(H) → M(K) for K ≤ H, ind^H_K : M(K) → M(H), and transfer (or conjugation) tr^g : M(H) → M(^g H) for g ∈ G, satisfying axioms such as res^G_H ∘ ind^G_K = ind^H_K ∘ res^K_H (projection formula), decompositions (Mackey formula), and Frobenius reciprocity. Mackey functors generalize the , where the category of R G-modules forms a Mackey functor via Hom and tensor operations. They form an Mack_R(G), with examples including the Burnside ring (counting orbits), groups H^n(G, U), and K-groups of group rings. This structure captures the dynamics of group actions on categories rather than sets, enabling computations in equivariant and . Mackey functors have applications in , where they classify equivariant spectra and compute groups, and in for decomposing induced modules via Mackey decomposition formulas. Historically, the concept was formalized in the building on George Mackey's work on induced representations, with key developments by Peter Webb and others in the for functorial approaches to .

References

  1. [1]
    Group Action -- from Wolfram MathWorld
    Subject classifications · Algebra · Group Theory · Group Operations.
  2. [2]
    [PDF] group actions - keith conrad
    All of our applications of group actions to group theory will flow from the relations between orbits, stabilizers, and fixed points, which we now make explicit ...
  3. [3]
    [PDF] GROUPS ACTING ON A SET 1. Left group actions Definition 1.1 ...
    A left (group) action of G on S is a rule for combining elements g ∈ G and elements x ∈ S, denoted by g.x. We additionally require the following 3 axioms. (0) ...
  4. [4]
    [PDF] Applications of Group Actions - MIT Mathematics
    Nov 25, 2019 · To do this, we begin with an introduction to group theory, developing the necessary tools we need to interrogate group actions. We begin by ...
  5. [5]
    [PDF] Chapter 9: Group actions - Mathematical and Statistical Sciences
    There is a rich theory of group actions, and it can be used to prove many deep results in group theory. M. Macauley (Clemson). Chapter 9: Group actions. Math ...
  6. [6]
    [PDF] 1. Group actions and other topics in group theory
    Oct 11, 2014 · In that context a. “topological group action” means a group action such that the action map G × X−→X is continuous (G × X has the product ...
  7. [7]
  8. [8]
    [PDF] Lecture #14 of 24 ∼ October 19th, 2020 - Math 5111 (Algebra 1)
    Oct 19, 2020 · There is also a notion of a right group action, which is a function from A × G to A whose [A1] statement is (a · g2) · g1 = a · (g2g1) and ...
  9. [9]
    [PDF] ADDITIONAL TOPICS IN GROUP THEORY 1. Order in Abelian ...
    A right action of a group G on a set X is a function ... Though part (c) indicates that there is no major difference between left actions and right actions,.
  10. [10]
    Transitive Group Action -- from Wolfram MathWorld
    A group action is transitive if it possesses only a single group orbit, i.e., for every pair of elements and , there is a group element such that . In this ...Missing: "abstract textbook
  11. [11]
    alternative characterization of multiply transitive permutation groups
    Mar 22, 2013 · Finally, note that the most common cases of n n -transitivity are for n=1 n = 1 (transitive), and n=2 n = 2 (doubly transitive). Title ...
  12. [12]
    doubly transitive groups are primitive - PlanetMath.org
    Mar 22, 2013 · Let G G acting on X X be doubly transitive. To show the action is , we must show that all blocks are trivial blocks; to do this, ...
  13. [13]
    Faithful Group Action -- from Wolfram MathWorld
    Subject classifications · Algebra · Group Theory · Group Operations.<|control11|><|separator|>
  14. [14]
    regular group action - PlanetMath.org
    Mar 22, 2013 · regular group action ... on a set X X . The action is called if for any pair α,β∈X α , β ∈ X there exists exactly one g∈G g ∈ G such that g⋅α=β g ...
  15. [15]
    blocks of permutation groups - PlanetMath
    Mar 22, 2013 · A block is a subset B B of A A such that for each σ∈G σ ∈ G , either σ⋅B=B σ ⋅ B = B or (σ⋅B)∩B=∅ ( σ ⋅ B ) ∩ B = ∅ , where σ⋅B={σ(b)∣b∈B} σ ...
  16. [16]
    Primitive Group Action -- from Wolfram MathWorld
    A primitive group action is transitive and it has no nontrivial group blocks. A transitive group action that is not primitive is called imprimitive.
  17. [17]
    [PDF] Primitive permutation groups 1 The basics 2 Minimal normal ...
    Aug 27, 2004 · Dixon and Brian Mortimer, Permutation Groups, Springer, 1996. [5] D. Gorenstein, Finite Simple Groups: An Introduction to their Classification,.
  18. [18]
    The O'Nan-Scott Theorem for Finite Primitive ... - bac-lac.gc.ca
    course, it is not a primitive action since Gα is not ... Rotman, An introduction to the theory of groups ... Suzuki, Group theory I (Springer, Berlin ...
  19. [19]
    [PDF] 5. Primitivity and related notions In this section we study some key ...
    Proposition 5.8 implies that, to understand the subgroup structure of Sym(n), we need to understand the finite primitive actions. Page 7. FINITE PERMUTATION ...
  20. [20]
    Group Orbit -- from Wolfram MathWorld
    The group orbit of a group element x can be defined as G(x)={gx in X:g in G}, where g runs over all elements of the group G.
  21. [21]
    orbit in nLab
    The category of orbits of a group G G is the full subcategory of the category of sets with an action of G G . Since any orbit of G G is isomorphic to the orbit ...Definition · Discrete case · Category of orbits · Topological case
  22. [22]
    invariant set in nLab
    Jan 31, 2025 · An invariant set is a subset which is invariant under a given action of a group or a monoid. 2. Definition
  23. [23]
    [PDF] GROUP ACTIONS ON SETS
    We first define this notion and give some examples. The structure of an action can be understood by means of orbits and stabilisers.
  24. [24]
    [PDF] Lecture 5.2: The orbit-stabilizer theorem
    If we have instead, a left group action, the proof carries through but using left cosets. M. Macauley (Clemson). Lecture 5.2: The orbit-stabilizer theorem. Math ...
  25. [25]
    Kombinatorische Anzahlbestimmungen für Gruppen, Graphen und ...
    Download PDF · Acta Mathematica ... Cite this article. Pólya, G. Kombinatorische Anzahlbestimmungen für Gruppen, Graphen und chemische Verbindungen.Missing: url | Show results with:url
  26. [26]
    [PDF] Roots of unity and cyclic groups - Williams College
    Among other results, we prove that a finite group is cyclic iff it doesn't have too many roots of unity. 1.
  27. [27]
    [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.
  28. [28]
    [PDF] On transitive and primitive dihedral groups of degree 2 (r≥2)
    Let G be a dihedral group of any order, then G is transitive. Proof. For given αi, αj as any two vertices of the regular polygon with i < j, we readily see that ...
  29. [29]
    [PDF] 19 Group Actions on G - 19.1 Conjugation - MIT OpenCourseWare
    19.1 Conjugation. Today, we will discuss the special case of group actions where the set S is G itself. We've seen the power of studying orbits and ...
  30. [30]
    [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.
  31. [31]
    Theory of Groups of Finite Order
    The British mathematician William Burnside (1852–1927) and Ferdinand Georg Frobenius (1849–1917), Professor at Zurich and Berlin universities, ...
  32. [32]
    [PDF] Counting colorful necklaces and bracelets in three colors
    Nov 26, 2018 · Group action, Burnside's lemma, Necklace, Bracelet,. Periodic three color sequences. 1. Introduction. A necklace with n beads and c colors is ...
  33. [33]
    [PDF] Counting Symmetries with Group Actions | MIT ESP
    Specifically, we introduce Burnside's Lemma, a tool that lets us count configurations of geometric figures that are preserved under symmetry. 1 What is a group?
  34. [34]
    [PDF] Counting List Colorings of Unlabeled Graphs - arXiv
    Sep 9, 2024 · In this paper, we pursue a different perspective and consider the problem of counting list colorings of unlabeled graphs. Unlike previous works, ...<|control11|><|separator|>
  35. [35]
    [PDF] Theory of Groups of Finite Order - Project Gutenberg
    The theory of groups of finite order may be said to date from the time of Cauchy. To him are due the first attempts at classification with a view to.
  36. [36]
    [PDF] 15 Permutation representations and G-sets
    For a given group G the collection of G-sets and G-equivariant map forms a category SetG. 15.8 Proposition. A G-equivariant map f : S → T is an isomorphism ...
  37. [37]
  38. [38]
    [PDF] group actions or permutation representations - People
    Thus a permutation representation of G on a set X is simply a group homomorphism from G to the group S(X) of all permutations of X. There is another point of ...
  39. [39]
    [PDF] be a group and X a non-empty set. A (right) group action of G on X is
    A (right) group action of G on X is a map. X × G → X, (x, g) 7→ x • g, such ... (2) The conjugation action of G on itself is defined by x • g := xg := g−1xg for.
  40. [40]
    [PDF] 1. Introduction Definition 1.1. Suppose G is a group. A G-set is a pair ...
    1.17. Suppose X and Y are two G-sets. An isomorphism of G-sets from X to Y is a mor- phism α : X → Y of G-sets which is one-one and onto. By Proposition 1.16, ...
  41. [41]
    [PDF] Group theory - UT Math
    This is just taking πV (g) := 1 for all g ∈ G, so every element of the group acts as the trivial automorphism on V . ... As such, this conjugate action is ...
  42. [42]
    [PDF] INTRODUCTION GROUP THEORY
    An isomorphism G →G is called an automorphism of G. 3.1.3. Remark. An ... (c) 'Conjugate action' defined g · h := ghg−1 for all g ∈ G, h ∈ X = G ...
  43. [43]
    [PDF] Lecture 1.2: Group actions - Mathematical and Statistical Sciences
    We say that “G acts on itself by right-multiplication.” M. Macauley (Clemson). Lecture 1.2: Group actions. Math 8510, Abstract Algebra I. 11 / 29. Page 12 ...
  44. [44]
    [PDF] Groups acting on themselves by left multiplication Conjugacy Classes
    We consider the action of G on itself via conjugation. Let G be a group and ... (h)|. We have seen that the orbits partition the group into classes, which we will ...
  45. [45]
    Definition of topological group acting on a topological space
    Mar 22, 2015 · The definition of a topological group G acting on a topological space X is there exists a continuous map from G×X→X such that eG.x=x for all x∈X ...
  46. [46]
    A continuous group action is an action by homeomorphisms
    Jan 3, 2022 · Lee states that, given a topological group G and a topological space X, an action is continuous if f:G×X→X is continuous, and the action is an action by ...
  47. [47]
    A reference to the fact that a topologically transitive action of a group ...
    Apr 4, 2020 · An action of a group G on a nonempty compact metrizable space K is topologically transitive (= the orbit GU of any nonempty open set U is dense) ...
  48. [48]
    [PDF] Variations on the Concept of Topological Transitivity - arXiv
    Jan 21, 2016 · This is equivalent to saying that the orbit O(x) = {fn(x) : n ∈ N} is dense in X. We denote by T rans(f) the set of transitive points.
  49. [49]
    [PDF] Proper Actions and Representation Theory - arXiv
    Definition–Lemma 3.4 (Proper Action). Let X be a locally compact topological space, on which a locally compact group G acts continu- ously. Then the ...
  50. [50]
    Why is the definition of a proper group action the way it is?
    Sep 18, 2017 · Let G be a topological group acting continuously on a topological space X. This means that G×X→X is a continuous function. A continuous map ...<|control11|><|separator|>
  51. [51]
    Proper Group Action -- from Wolfram MathWorld
    Proper Group Action ... is a proper map, i.e., inverses of compact sets are compact. A proper action must have compact isotropy groups at all points of X . See ...Missing: definition | Show results with:definition
  52. [52]
    [PDF] Representations of SO(3) in C [x, y, z] - kth .diva
    We define SO(3) as the group of all the possible rotations in 3-dimensional space R3 about the origin. ... Since SO(3) is a compact topological group, we know ...<|control11|><|separator|>
  53. [53]
    Linear Representations of Finite Groups - SpringerLink
    Book Title: Linear Representations of Finite Groups · Authors: Jean-Pierre Serre · Series Title: Graduate Texts in Mathematics · Publisher: Springer New York, NY.
  54. [54]
    [PDF] FROM GROUPS TO GROUPOIDS: A BRIEF SURVEY - Ronald Brown
    A groupoid is like a group with many objects, or many identities. A groupoid with one object is essentially a group.
  55. [55]
    [PDF] arXiv:math/0302182v1 [math.AT] 15 Feb 2003
    In section 5, to make certain reductions in the presentation problem, we will need the notion of a group action on a groupoid and the resulting semidirect ...
  56. [56]
    [PDF] The groupoid structure of groupoid morphisms - UC Berkeley math
    Aug 12, 2019 · In this paper we study morphisms and automorphisms of groupoids. Our motivation comes from the study of maps between orbifolds and group ...
  57. [57]
    [1410.6981] Pseudogroups via pseudoactions: Unifying local, global ...
    Oct 26, 2014 · A pseudoaction generates a pseudogroup of transformations of M in the same way an ordinary Lie group action generates a transformation group.
  58. [58]
    [0812.4864] Groupoidification Made Easy - arXiv
    Dec 30, 2008 · Groupoidification is a form of categorification in which vector spaces are replaced by groupoids, and linear operators are replaced by spans of groupoids.Missing: group actions
  59. [59]
    Partial Actions of Groups and Actions of Inverse Semigroups on ...
    Ruy Exel, Partial Actions of Groups and Actions of Inverse Semigroups, Proceedings of the American Mathematical Society, Vol. 126, No. 12 (Dec., 1998), pp.
  60. [60]
    [PDF] arXiv:2102.06723v1 [math.RA] 12 Feb 2021
    Feb 12, 2021 · Upcoming work of the author computes the homotopy groups of this spec- trum as Mackey functors of K-groups of twisted group rings [Bra21], ...
  61. [61]
    [PDF] C*-algebras and Mackey's theory of group representations
    The aim of the "Mackey machine" is to describe the representation theory of such a crossed product in terms of knowledge of C· (N) and of the action of G on it ...
  62. [62]
    [PDF] A Guide to Mackey Functors
    The automorphisms of each group G act on M(G), and because the inner automorphisms act trivially each. M(G) has the structure of an R[OutG]-module. The main ...
  63. [63]
    [PDF] Permutation modules, Mackey functors, and Artin motives
    Continuity of the action on an R-module M is equivalent to every stabilizer sub- group Γm := γ ∈ Γ γ · m = m being open. The tensor product over R with the ...
  64. [64]
    [1006.4975] Crossed products and the Mackey-Rieffel-Green machine
    Jun 25, 2010 · We give an introduction into the ideal structure and representation theory of crossed products by actions of locally compact groups on C*- ...
  65. [65]
    [PDF] George W. Mackey - Biographical Memoirs
    Some facts about actions of locally compact groups on Borel spaces and measure spaces constituted another building block for the Mackey machine. A group ...
  66. [66]
    [PDF] George Mackey and His Work on Representation Theory and ...
    His ideas in ergodic theory and the ergodic aspects of group actions were the first sightings of a huge new continent that Connes explored later—the theory of ...