Jump to content
 







Main menu
   


Navigation  



Main page
Contents
Current events
Random article
About Wikipedia
Contact us
Donate
 




Contribute  



Help
Learn to edit
Community portal
Recent changes
Upload file
 








Search  

































Create account

Log in
 









Create account
 Log in
 




Pages for logged out editors learn more  



Contributions
Talk
 



















Contents

   



(Top)
 


1 Definition  



1.1  Left group action  





1.2  Right group action  







2 Notable properties of actions  



2.1  Transitivity properties  



2.1.1  Examples  







2.2  Primitive actions  





2.3  Topological properties  





2.4  Actions of topological groups  





2.5  Linear actions  







3 Orbits and stabilizers  



3.1  Invariant subsets  





3.2  Fixed points and stabilizer subgroups  





3.3  Orbit-stabilizer theorem and Burnside's lemma  







4 Examples  





5 Group actions and groupoids  





6 Morphisms and isomorphisms between G-sets  





7 Variants and generalizations  





8 Gallery  





9 See also  





10 Notes  





11 Citations  





12 References  





13 External links  














Group action






Català
Čeština
Deutsch
Eesti
Español
Esperanto
فارسی
Français

Bahasa Indonesia
Italiano
עברית
Magyar
Nederlands

Piemontèis
Polski
Português
Русский
Svenska
Українська
Tiếng Vit


 

Edit links
 









Article
Talk
 

















Read
Edit
View history
 








Tools
   


Actions  



Read
Edit
View history
 




General  



What links here
Related changes
Upload file
Special pages
Permanent link
Page information
Cite this page
Get shortened URL
Download QR code
Wikidata item
 




Print/export  



Download as PDF
Printable version
 




In other projects  



Wikimedia Commons
Wikibooks
 
















Appearance
   

 






From Wikipedia, the free encyclopedia
 

(Redirected from Trivial action)

The cyclic group C3 consisting of the rotations by 0°, 120° and 240° acts on the set of the three vertices.

Inmathematics, many sets of transformations form a group under function composition; for example, the rotations around a point in the plane. It is often useful to consider the group as an abstract group, and to say that one has a group action of the abstract group that consists of performing the transformations of the group of transformations. The reason for distinguishing the group from the transformations is that, generally, a group of transformations of a structure acts also on various related structures; for example, the above rotation group acts also on triangles by transforming triangles into triangles.

Formally, a group action of a group G on a set S is a group homomorphism from G to some group (under function composition) of functions from S to itself.

If a group acts on a structure, it will usually also act on objects built from that structure. For example, the group of Euclidean isometries acts on Euclidean space and also on the figures drawn in it; in particular, it acts on the set of all triangles. Similarly, the group of symmetries of a polyhedron acts on the vertices, the edges, and the faces of the polyhedron.

A group action on a vector space is called a representation of the group. In the case of a finite-dimensional vector space, it allows one to identify many groups with subgroups of the general linear group GL(n, K), the group of the invertible matricesofdimension n over a field K.

The symmetric group Sn acts on any set with n elements by permuting the elements of the set. Although the group of all permutations of a set depends formally on the set, the concept of group action allows one to consider a single group for studying the permutations of all sets with the same cardinality.

Definition[edit]

Left group action[edit]

IfG is a group with identity element e, and X is a set, then a (left) group action αofGonX is a function

that satisfies the following two axioms:[1]

Identity:
Compatibility:

for all g and hinG and all xinX.

The group G is then said to act on X (from the left). A set X together with an action of G is called a (left) G-set.

It can be notationally convenient to curry the action α, so that, instead, one has a collection of transformations αg : XX, with one transformation αg for each group element gG. The identity and compatibility relations then read

and

with being function composition. The second axiom then states that the function composition is compatible with the group multiplication; they form a commutative diagram. This axiom can be shortened even further, and written as αgαh = αgh.

With the above understanding, it is very common to avoid writing α entirely, and to replace it with either a dot, or with nothing at all. Thus, α(g, x) can be shortened to gxorgx, especially when the action is clear from context. The axioms are then

From these two axioms, it follows that for any fixed ginG, the function from X to itself which maps xtogx is a bijection, with inverse bijection the corresponding map for g−1. Therefore, one may equivalently define a group action of GonX as a group homomorphism from G into the symmetric group Sym(X) of all bijections from X to itself.[2]

Right group action[edit]

Likewise, a right group actionofGonX is a function

that satisfies the analogous axioms:[3]

Identity:
Compatibility:

(with α(x, g) often shortened to xgorxg when the action being considered is clear from context)

Identity:
Compatibility:

for all g and hinG and all xinX.

The difference between left and right actions is in the order in which a product gh acts on x. For a left action, h acts first, followed by g second. For a right action, g acts first, followed by h second. Because of the formula (gh)−1 = h−1g−1, a left action can be constructed from a right action by composing with the inverse operation of the group. Also, a right action of a group GonX can be considered as a left action of its opposite group GoponX.

Thus, for establishing general properties of group actions, it suffices to consider only left actions. However, there are cases where this is not possible. For example, the multiplication of a group induces both a left action and a right action on the group itself—multiplication on the left and on the right, respectively.

Notable properties of actions[edit]

Let G be a group acting on a set X. The action is called faithfuloreffectiveifgx = x for all xX implies that g = eG. Equivalently, the homomorphism from G to the group of bijections of X corresponding to the action is injective.

The action is called free (orsemiregularorfixed-point free) if the statement that gx = x for some xX already implies that g = eG. In other words, no non-trivial element of G fixes a point of X. This is a much stronger property than faithfulness.

For example, the action of any group on itself by left multiplication is free. This observation implies Cayley's theorem that any group can be embedded in a symmetric group (which is infinite when the group is). A finite group may act faithfully on a set of size much smaller than its cardinality (however such an action cannot be free). For instance the abelian 2-group (Z / 2Z)n (of cardinality 2n) acts faithfully on a set of size 2n. This is not always the case, for example the cyclic group Z / 2nZ cannot act faithfully on a set of size less than 2n.

In general the smallest set on which a faithful action can be defined can vary greatly for groups of the same size. For example, three groups of size 120 are the symmetric group S5, the icosahedral group A5 × Z / 2Z and the cyclic group Z / 120Z. The smallest sets on which faithful actions can be defined for these groups are of size 5, 7, and 16 respectively.

Transitivity properties[edit]

The action of GonX is called transitive if for any two points x, yX there exists a gG so that gx = y.

The action is simply transitive (orsharply transitive, or regular) if it is both transitive and free. This means that given x, yX the element g in the definition of transitivity is unique. If X is acted upon simply transitively by a group G then it is called a principal homogeneous space for G or a G-torsor.

For an integer n ≥ 1, the action is n-transitiveifX has at least n elements, and for any pair of n-tuples (x1, ..., xn), (y1, ..., yn) ∈ Xn with pairwise distinct entries (that is xixj, yiyj when ij) there exists a gG such that gxi = yi for i = 1, ..., n. In other words the action on the subset of Xn of tuples without repeated entries is transitive. For n = 2, 3 this is often called double, respectively triple, transitivity. The class of 2-transitive groups (that is, subgroups of a finite symmetric group whose action is 2-transitive) and more generally multiply transitive groups is well-studied in finite group theory.

An action is sharply n-transitive when the action on tuples without repeated entries in Xn is sharply transitive.

Examples[edit]

The action of the symmetric group of X is transitive, in fact n-transitive for any n up to the cardinality of X. If X has cardinality n, the action of the alternating groupis(n − 2)-transitive but not (n − 1)-transitive.

The action of the general linear group of a vector space V on the set V ∖ {0} of non-zero vectors is transitive, but not 2-transitive (similarly for the action of the special linear group if the dimension of v is at least 2). The action of the orthogonal group of a Euclidean space is not transitive on nonzero vectors but it is on the unit sphere.

Primitive actions[edit]

The action of GonX is called primitive if there is no partitionofX preserved by all elements of G apart from the trivial partitions (the partition in a single piece and its dual, the partition into singletons).

Topological properties[edit]

Assume that X is a topological space and the action of G is by homeomorphisms.

The action is wandering if every xX has a neighbourhood U such that there are only finitely many gG with gUU ≠ ∅.[4]

More generally, a point xX is called a point of discontinuity for the action of G if there is an open subset Ux such that there are only finitely many gG with gUU ≠ ∅. The domain of discontinuity of the action is the set of all points of discontinuity. Equivalently it is the largest G-stable open subset Ω ⊂ X such that the action of GonΩ is wandering.[5] In a dynamical context this is also called a wandering set.

The action is properly discontinuous if for every compact subset KX there are only finitely many gG such that gKK ≠ ∅. This is strictly stronger than wandering; for instance the action of ZonR2 ∖ {(0, 0)} given by n⋅(x, y) = (2nx, 2ny) is wandering and free but not properly discontinuous.[6]

The action by deck transformations of the fundamental group of a locally simply connected space on an covering space is wandering and free. Such actions can be characterized by the following property: every xX has a neighbourhood U such that gUU = ∅ for every gG ∖ {eG}.[7] Actions with this property are sometimes called freely discontinuous, and the largest subset on which the action is freely discontinuous is then called the free regular set.[8]

An action of a group G on a locally compact space X is called cocompact if there exists a compact subset AX such that X = GA. For a properly discontinuous action, cocompactness is equivalent to compactness of the quotient space G \ X.

Actions of topological groups[edit]

Now assume G is a topological group and X a topological space on which it acts by homeomorphisms. The action is said to be continuous if the map G × XX is continuous for the product topology.

The action is said to be proper if the map G × XX × X defined by (g, x) ↦ (x, gx)isproper.[9] This means that given compact sets K, K the set of gG such that gKK′ ≠ ∅ is compact. In particular, this is equivalent to proper discontinuity G is a discrete group.

It is said to be locally free if there exists a neighbourhood UofeG such that gxx for all xX and gU ∖ {eG}.

The action is said to be strongly continuous if the orbital map ggx is continuous for every xX. Contrary to what the name suggests, this is a weaker property than continuity of the action.[citation needed]

IfG is a Lie group and Xadifferentiable manifold, then the subspace of smooth points for the action is the set of points xX such that the map ggxissmooth. There is a well-developed theory of Lie group actions, i.e. action which are smooth on the whole space.

Linear actions[edit]

Ifg acts by linear transformations on a module over a commutative ring, the action is said to be irreducible if there are no proper nonzero g-invariant submodules. It is said to be semisimple if it decomposes as a direct sum of irreducible actions.

Orbits and stabilizers[edit]

In the compound of five tetrahedra, the symmetry group is the (rotational) icosahedral group I of order 60, while the stabilizer of a single chosen tetrahedron is the (rotational) tetrahedral group T of order 12, and the orbit space I / T (of order 60/12 = 5) is naturally identified with the 5 tetrahedra – the coset gT corresponds to the tetrahedron to which g sends the chosen tetrahedron.

Consider a group G acting on a set X. The orbit of an element xinX is the set of elements in X to which x can be moved by the elements of G. The orbit of x is denoted by Gx:

The defining properties of a group guarantee that the set of orbits of (points xin) X under the action of G form a partitionofX. The associated equivalence relation is defined by saying x ~ y if and only if there exists a ginG with gx = y. The orbits are then the equivalence classes under this relation; two elements x and y are equivalent if and only if their orbits are the same, that is, Gx = Gy.

The group action is transitive if and only if it has exactly one orbit, that is, if there exists xinX with Gx = X. This is the case if and only if Gx = X for all xinX (given that X is non-empty).

The set of all orbits of X under the action of G is written as X / G (or, less frequently, as G \ X), and is called the quotient of the action. In geometric situations it may be called the orbit space, while in algebraic situations it may be called the space of coinvariants, and written XG, by contrast with the invariants (fixed points), denoted XG: the coinvariants are a quotient while the invariants are a subset. The coinvariant terminology and notation are used particularly in group cohomology and group homology, which use the same superscript/subscript convention.

Invariant subsets[edit]

IfY is a subsetofX, then GY denotes the set {gy : gG and yY}. The subset Y is said to be invariant under GifGY = Y (which is equivalent GYY). In that case, G also operates on Yby restricting the action to Y. The subset Y is called fixed under Gifgy = y for all ginG and all yinY. Every subset that is fixed under G is also invariant under G, but not conversely.

Every orbit is an invariant subset of X on which G acts transitively. Conversely, any invariant subset of X is a union of orbits. The action of GonXistransitive if and only if all elements are equivalent, meaning that there is only one orbit.

AG-invariant element of XisxX such that gx = x for all gG. The set of all such x is denoted XG and called the G-invariantsofX. When X is a G-module, XG is the zeroth cohomology group of G with coefficients in X, and the higher cohomology groups are the derived functors of the functorofG-invariants.

Fixed points and stabilizer subgroups[edit]

Given ginG and xinX with gx = x, it is said that "x is a fixed point of g" or that "g fixes x". For every xinX, the stabilizer subgroupofG with respect to x (also called the isotropy grouporlittle group[10]) is the set of all elements in G that fix x: This is a subgroupofG, though typically not a normal one. The action of GonXisfree if and only if all stabilizers are trivial. The kernel N of the homomorphism with the symmetric group, G → Sym(X), is given by the intersection of the stabilizers Gx for all xinX. If N is trivial, the action is said to be faithful (or effective).

Let x and y be two elements in X, and let g be a group element such that y = gx. Then the two stabilizer groups Gx and Gy are related by Gy = gGxg−1. Proof: by definition, hGy if and only if h⋅(gx) = gx. Applying g−1 to both sides of this equality yields (g−1hg)⋅x = x; that is, g−1hgGx. An opposite inclusion follows similarly by taking hGx and x = g−1y.

The above says that the stabilizers of elements in the same orbit are conjugate to each other. Thus, to each orbit, we can associate a conjugacy class of a subgroup of G (that is, the set of all conjugates of the subgroup). Let (H) denote the conjugacy class of H. Then the orbit O has type (H) if the stabilizer Gx of some/any xinO belongs to (H). A maximal orbit type is often called a principal orbit type.

Orbit-stabilizer theorem and Burnside's lemma[edit]

Orbits and stabilizers are closely related. For a fixed xinX, consider the map f : GX given by ggx. By definition the image f(G) of this map is the orbit Gx. The condition for two elements to have the same image is In other words, f(g) = f(h) if and only if g and h lie in the same coset for the stabilizer subgroup Gx. Thus, the fiber f−1({y})off over any yinGx is contained in such a coset, and every such coset also occurs as a fiber. Therefore f induces a bijection between the set G / Gx of cosets for the stabilizer subgroup and the orbit Gx, which sends gGxgx.[11] This result is known as the orbit-stabilizer theorem.

IfG is finite then the orbit-stabilizer theorem, together with Lagrange's theorem, gives in other words the length of the orbit of x times the order of its stabilizer is the order of the group. In particular that implies that the orbit length is a divisor of the group order.

Example: Let G be a group of prime order p acting on a set X with k elements. Since each orbit has either 1orp elements, there are at most k mod p orbits of length 1 which are G-invariant elements.

This result is especially useful since it can be employed for counting arguments (typically in situations where X is finite as well).

Cubical graph with vertices labeled
Example: We can use the orbit-stabilizer theorem to count the automorphisms of a graph. Consider the cubical graph as pictured, and let G denote its automorphism group. Then G acts on the set of vertices {1, 2, ..., 8}, and this action is transitive as can be seen by composing rotations about the center of the cube. Thus, by the orbit-stabilizer theorem, |G| = |G ⋅ 1| |G1| = 8 |G1|. Applying the theorem now to the stabilizer G1, we can obtain |G1| = |(G1) ⋅ 2| |(G1)2|. Any element of G that fixes 1 must send 2 to either 2, 4, or 5. As an example of such automorphisms consider the rotation around the diagonal axis through 1 and 7 by 2π/3, which permutes 2, 4, 5 and 3, 6, 8, and fixes 1 and 7. Thus, |(G1) ⋅ 2| = 3. Applying the theorem a third time gives |(G1)2| = |((G1)2) ⋅ 3| |((G1)2)3|. Any element of G that fixes 1 and 2 must send 3 to either 3 or 6. Reflecting the cube at the plane through 1, 2, 7 and 8 is such an automorphism sending 3 to 6, thus |((G1)2) ⋅ 3| = 2. One also sees that ((G1)2)3 consists only of the identity automorphism, as any element of G fixing 1, 2 and 3 must also fix all other vertices, since they are determined by their adjacency to 1, 2 and 3. Combining the preceding calculations, we can now obtain |G| = 8 ⋅ 3 ⋅ 2 ⋅ 1 = 48.

A result closely related to the orbit-stabilizer theorem is Burnside's lemma: where Xg is the set of points fixed by g. This result is mainly of use when G and X are finite, when it can be interpreted as follows: the number of orbits is equal to the average number of points fixed per group element.

Fixing a group G, the set of formal differences of finite G-sets forms a ring called the Burnside ringofG, where addition corresponds to disjoint union, and multiplication to Cartesian product.

Examples[edit]

Group actions and groupoids[edit]

The notion of group action can be encoded by the action groupoid G′ = GX associated to the group action. The stabilizers of the action are the vertex groups of the groupoid and the orbits of the action are its components.

Morphisms and isomorphisms between G-sets[edit]

IfX and Y are two G-sets, a morphism from XtoY is a function f : XY such that f(gx) = gf(x) for all ginG and all xinX. Morphisms of G-sets are also called equivariant mapsorG-maps.

The composition of two morphisms is again a morphism. If a morphism f is bijective, then its inverse is also a morphism. In this case f is called an isomorphism, and the two G-sets X and Y are called isomorphic; for all practical purposes, isomorphic G-sets are indistinguishable.

Some example isomorphisms:

With this notion of morphism, the collection of all G-sets forms a category; this category is a Grothendieck topos (in fact, assuming a classical metalogic, this topos will even be Boolean).

Variants and generalizations[edit]

We can also consider actions of monoids on sets, by using the same two axioms as above. This does not define bijective maps and equivalence relations however. See semigroup action.

Instead of actions on sets, we can define actions of groups and monoids on objects of an arbitrary category: start with an object X of some category, and then define an action on X as a monoid homomorphism into the monoid of endomorphismsofX. If X has an underlying set, then all definitions and facts stated above can be carried over. For example, if we take the category of vector spaces, we obtain group representations in this fashion.

We can view a group G as a category with a single object in which every morphism is invertible.[14] A (left) group action is then nothing but a (covariant) functor from G to the category of sets, and a group representation is a functor from G to the category of vector spaces.[15] A morphism between G-sets is then a natural transformation between the group action functors.[16] In analogy, an action of a groupoid is a functor from the groupoid to the category of sets or to some other category.

In addition to continuous actions of topological groups on topological spaces, one also often considers smooth actions of Lie groups on smooth manifolds, regular actions of algebraic groupsonalgebraic varieties, and actionsofgroup schemesonschemes. All of these are examples of group objects acting on objects of their respective category.

Gallery[edit]

See also[edit]

Notes[edit]

Citations[edit]

  1. ^ Eie & Chang (2010). A Course on Abstract Algebra. p. 144.
  • ^ This is done, for example, by Smith (2008). Introduction to abstract algebra. p. 253.
  • ^ "Definition:Right Group Action Axioms". Proof Wiki. Retrieved 19 December 2021.
  • ^ Thurston 1997, Definition 3.5.1(iv).
  • ^ Kapovich 2009, p. 73.
  • ^ Thurston 1980, p. 176.
  • ^ Hatcher 2002, p. 72.
  • ^ Maskit 1988, II.A.1, II.A.2.
  • ^ tom Dieck 1987.
  • ^ Procesi, Claudio (2007). Lie Groups: An Approach through Invariants and Representations. Springer Science & Business Media. p. 5. ISBN 9780387289298. Retrieved 23 February 2017.
  • ^ M. Artin, Algebra, Proposition 6.8.4 on p. 179
  • ^ Eie & Chang (2010). A Course on Abstract Algebra. p. 145.
  • ^ Reid, Miles (2005). Geometry and topology. Cambridge, UK New York: Cambridge University Press. p. 170. ISBN 9780521613255.
  • ^ Perrone (2024), pp. 7–9
  • ^ Perrone (2024), pp. 36–39
  • ^ Perrone (2024), pp. 69–71
  • References[edit]

    External links[edit]


    Retrieved from "https://en.wikipedia.org/w/index.php?title=Group_action&oldid=1231272744#trivial"

    Categories: 
    Group theory
    Group actions (mathematics)
    Representation theory of groups
    Symmetry
    Hidden categories: 
    Articles with short description
    Short description is different from Wikidata
    All articles with unsourced statements
    Articles with unsourced statements from May 2023
    All accuracy disputes
    Articles with disputed statements from March 2015
    Articles with J9U identifiers
    Articles with LCCN identifiers
     



    This page was last edited on 27 June 2024, at 12:41 (UTC).

    Text is available under the Creative Commons Attribution-ShareAlike License 4.0; additional terms may apply. By using this site, you agree to the Terms of Use and Privacy Policy. Wikipedia® is a registered trademark of the Wikimedia Foundation, Inc., a non-profit organization.



    Privacy policy

    About Wikipedia

    Disclaimers

    Contact Wikipedia

    Code of Conduct

    Developers

    Statistics

    Cookie statement

    Mobile view



    Wikimedia Foundation
    Powered by MediaWiki