Topological group


In mathematics, a topological group is a group together with a topology on such that both the group's binary operation and the function mapping group elements to their respective inverses are continuous functions with respect to the topology.
A topological group is a mathematical object with both an algebraic structure and a topological structure.
Thus, one may perform algebraic operations, because of the group structure, and one may talk about continuous functions, because of the topology.
Topological groups, along with continuous group actions, are used to study continuous symmetries, which have many applications, for example, in physics.

Formal definition

A topological group,, is a topological space that is also a group such that the group operation :
and inversion map:
are continuous
Here is viewed as a topological space with the product topology.
Such a topology is said to be compatible with the group operations and is called a group topology.
;Checking continuity
The product map is continuous if and only if for any and any neighborhood of in, there exist neighborhoods of and of in such that, where.
The inversion map is continuous if and only if for any and any neighborhood of in, there exist neighborhoods of in such that, where.
To show that a topology is compatible with the group operations, it suffices to check that the map
is continuous.
Explicitly, this means that for any and any neighborhood in of, there exist neighborhoods of and of in such that.
;Additive notation
This definition used notation for multiplicative groups;
the equivalent for additive groups would be that the following two operations are continuous:
;Hausdorffness
Although not part of this definition, many authors require that the topology on be Hausdorff.
One reason for this is that any topological group can be canonically associated with a Hausdorff topological group by taking an appropriate canonical quotient;
this however, often still requires working with the original non-Hausdorff topological group.
Other reasons, and some equivalent conditions, are discussed below.
This article will not assume that topological groups are necessarily Hausdorff.
;Category
In the language of category theory, topological groups can be defined concisely as group objects in the category of topological spaces, in the same way that ordinary groups are group objects in the category of sets.
Note that the axioms are given in terms of the maps, hence are categorical definitions.

Homomorphisms

A homomorphism of topological groups means a continuous group homomorphism.
Topological groups, together with their homomorphisms, form a category.
A group homomorphism between commutative topological groups is continuous if and only if it is continuous at some point.
An isomorphism of topological groups is a group isomorphism that is also a homeomorphism of the underlying topological spaces.
This is stronger than simply requiring a continuous group isomorphism—the inverse must also be continuous.
There are examples of topological groups that are isomorphic as ordinary groups but not as topological groups.
Indeed, any non-discrete topological group is also a topological group when considered with the discrete topology.
The underlying groups are the same, but as topological groups there is not an isomorphism.

Examples

Every group can be trivially made into a topological group by considering it with the discrete topology; such groups are called discrete groups.
In this sense, the theory of topological groups subsumes that of ordinary groups.
The indiscrete topology also makes every group into a topological group.
The real numbers, with the usual topology form a topological group under addition.
Euclidean -space is also a topological group under addition, and more generally, every topological vector space forms an topological group.
Some other examples of abelian topological groups are the circle group, or the torus for any natural number.
The classical groups are important examples of non-abelian topological groups. For instance, the general linear group of all invertible -by- matrices with real entries can be viewed as a topological group with the topology defined by viewing as a subspace of Euclidean space.
Another classical group is the orthogonal group, the group of all linear maps from to itself that preserve the length of all vectors.
The orthogonal group is compact as a topological space. Much of Euclidean geometry can be viewed as studying the structure of the orthogonal group, or the closely related group of isometries of.
The groups mentioned so far are all Lie groups, meaning that they are smooth manifolds in such a way that the group operations are smooth, not just continuous.
Lie groups are the best-understood topological groups; many questions about Lie groups can be converted to purely algebraic questions about Lie algebras and then solved.
An example of a topological group that is not a Lie group is the additive group of rational numbers, with the topology inherited from.
This is a countable space, and it does not have the discrete topology.
An important example for number theory is the group of p-adic integers, for a prime number, meaning the inverse limit of the finite groups as n goes to infinity.
The group is well behaved in that it is compact, but it differs from Lie groups in that it is totally disconnected.
More generally, there is a theory of p-adic Lie groups, including compact groups such as as well as locally compact groups such as, where is the locally compact field of p-adic numbers.
The group is a pro-finite group; it is isomorphic to a subgroup of the product in such a way that its topology is induced by the product topology, where the finite groups are given the discrete topology.
Another large class of pro-finite groups important in number theory are absolute Galois groups.
Some topological groups can be viewed as infinite dimensional Lie groups; this phrase is best understood informally, to include several different families of examples.
For example, a topological vector space, such as a Banach space or Hilbert space, is an abelian topological group under addition. Some other infinite-dimensional groups that have been studied, with varying degrees of success, are loop groups, Kac–Moody groups, diffeomorphism groups, homeomorphism groups, and gauge groups.
In every Banach algebra with multiplicative identity, the set of invertible elements forms a topological group under multiplication.
For example, the group of invertible bounded operators on a Hilbert space arises this way.

Properties

;Translation invariance
The inversion operation on a topological group is a homeomorphism from to itself.
Likewise, if a is any element of, then left or right multiplication by a yields a homeomorphism.
Consequently, if is a neighborhood basis of the identity element in a commutative topological group then for all,
is a neighborhood basis of in, where.
In particular, any group topology on a commutative topological group is completely determined by the any neighborhood basis at the identity element.
If is any subset of and is an open subset of, then is an open subset of.
;Symmetric neighborhoods
A subset of is said to be symmetric if.
The closure of every symmetric set in a commutative topological group is symmetric.
If is any subset of a commutative topological group, then the following sets are also symmetric:,, and.
For any neighborhood in a commutative topological group of the identity element, there exists a symmetric neighborhood of the identity element such that, where note that is necessarily a symmetric neighborhood of the identity element.
Thus every topological group has a neighborhood basis at the identity element consisting of symmetric sets.
If is a locally compact commutative group, then for any neighborhood in of the identity element, there exists a symmetric relatively compact neighborhood of the identity element such that .
;Uniform space
Every topological group can be viewed as a uniform space in two ways; the left uniformity turns all left multiplications into uniformly continuous maps while the right uniformity turns all right multiplications into uniformly continuous maps.
If is not abelian, then these two need not coincide.
The uniform structures allow one to talk about notions such as completeness, uniform continuity and uniform convergence on topological groups.
;Separation properties
If is an open subset of a commutative topological group and contains a compact set, then there exists a neighborhood of the identity element such that.
As a uniform space, every commutative topological group is completely regular.
Consequently, for a multiplicative topological group with identity element 1, the following are equivalent:

  1. is a T0-space ;
  2. is a T2-space ;
  3. is a T ;
  4. is closed in ;
  5. , where is a neighborhood basis of the identity element in ;
  6. for any such that, there exists a neighborhood in of the identity element such that.
A subgroup of a commutative topological group is discrete if and only if it has an isolated point.
If is not Hausdorff, then one can obtain a Hausdorff group by passing to the quotient group, where is the closure of the identity.
This is equivalent to taking the Kolmogorov quotient of.
;Metrizability
The Birkhoff–Kakutani theorem states that the following three conditions on a topological group are equivalent:
;Subgroups
Every subgroup of a topological group is itself a topological group when given the subspace topology.
Every open subgroup is also closed in, since the complement of is the open set given by the union of open sets for.
If is a subgroup of then the closure of is also a subgroup.
Likewise, if is a normal subgroup of, the closure of is normal in.
;Quotients and normal subgroups
If is a subgroup of, the set of left cosets with the quotient topology is called a homogeneous space for.
The quotient map is always open.
For example, for a positive integer, the sphere is a homogeneous space for the rotation group in, with.
A homogeneous space is Hausdorff if and only if is closed in.
Partly for this reason, it is natural to concentrate on closed subgroups when studying topological groups.
If is a normal subgroup of, then the quotient group becomes a topological group when given the quotient topology.
It is Hausdorff if and only if is closed in.
For example, the quotient group is isomorphic to the circle group.
In any topological group, the identity component is a closed normal subgroup.
If is the identity component and a is any point of, then the left coset is the component of containing a.
So the collection of all left cosets of in is equal to the collection of all components of.
It follows that the quotient group is totally disconnected.
;Closure and compactness
In any commutative topological group, the product of a compact set and a closed set is a closed set.
Furthermore, for any subsets and of,.
If is a subgroup of a commutative topological group and if is a neighborhood in of the identity element such that is closed, then is closed.
Every discrete subgroup of a Hausdorff commutative topological group is closed.
;Isomorphism theorems
The isomorphism theorems from ordinary group theory are not always true in the topological setting.
This is because a bijective homomorphism need not be an isomorphism of topological groups.
The theorems are valid if one places certain restrictions on the maps involved.
For example, the first isomorphism theorem states that if is a homomorphism, then the homomorphism from to mvar|f

Hilbert's fifth problem

There are several strong results on the relation between topological groups and Lie groups.
First, every continuous homomorphism of Lie groups is smooth.
It follows that a topological group has a unique structure of a Lie group if one exists.
Also, Cartan's theorem says that every closed subgroup of a Lie group is a Lie subgroup, in particular a smooth submanifold.
Hilbert's fifth problem asked whether a topological group that is a topological manifold must be a Lie group.
In other words, does have the structure of a smooth manifold, making the group operations smooth?
As shown by Andrew Gleason, Deane Montgomery, and Leo Zippin, the answer to this problem is yes.
In fact, has a real analytic structure.
Using the smooth structure, one can define the Lie algebra of, an object of linear algebra that determines a connected group up to covering spaces.
As a result, the solution to Hilbert's fifth problem reduces the classification of topological groups that are topological manifolds to an algebraic problem, albeit a complicated problem in general.
The theorem also has consequences for broader classes of topological groups. First, every compact group is an inverse limit of compact Lie groups.

Furthermore, every connected locally compact group is an inverse limit of connected Lie groups.
At the other extreme, a totally disconnected locally compact group always contains a compact open subgroup, which is necessarily a profinite group.

Representations of compact or locally compact groups

An action of a topological group on a topological space X is a group action of on X such that the corresponding function is continuous.
Likewise, a representation of a topological group on a real or complex topological vector space V is a continuous action of on V such that for each, the map from V to itself is linear.
Group actions and representation theory are particularly well understood for compact groups, generalizing what happens for finite groups.
For example, every finite-dimensional representation of a compact group is a direct sum of irreducible representations.
An infinite-dimensional unitary representation of a compact group can be decomposed as a Hilbert-space direct sum of irreducible representations, which are all finite-dimensional; this is part of the Peter–Weyl theorem.
For example, the theory of Fourier series describes the decomposition of the unitary representation of the circle group on the complex Hilbert space.
The irreducible representations of are all 1-dimensional, of the form for integers .
Each of these representations occurs with multiplicity 1 in.
The irreducible representations of all compact connected Lie groups have been classified.
In particular, the character of each irreducible representation is given by the Weyl character formula.
More generally, locally compact groups have a rich theory of harmonic analysis, because they admit a natural notion of measure and integral, given by the Haar measure.
Every unitary representation of a locally compact group can be described as a direct integral of irreducible unitary representations.

A basic example is the Fourier transform, which decomposes the action of the additive group on the Hilbert space as a direct integral of the irreducible unitary representations of.
The irreducible unitary representations of are all 1-dimensional, of the form for.
The irreducible unitary representations of a locally compact group may be infinite-dimensional.
A major goal of representation theory, related to the Langlands classification of admissible representations, is to find the unitary dual for the semisimple Lie groups.
The unitary dual is known in many cases such as representation theory of SL2|, but not all.
For a locally compact abelian group, every irreducible unitary representation has dimension 1.
In this case, the unitary dual is a group, in fact another locally compact abelian group.
Pontryagin duality states that for a locally compact abelian group, the dual of is the original group.
For example, the dual group of the integers is the circle group, while the group of real numbers is isomorphic to its own dual.
Every locally compact group has a good supply of irreducible unitary representations; for example, enough representations to distinguish the points of .
By contrast, representation theory for topological groups that are not locally compact has so far been developed only in special situations, and it may not be reasonable to expect a general theory.
For example, there are many abelian Banach–Lie groups for which every representation on Hilbert space is trivial.

Homotopy theory of topological groups

Topological groups are special among all topological spaces, even in terms of their homotopy type.
One basic point is that a topological group determines a path-connected topological space, the classifying space .
The group is isomorphic in the homotopy category to the loop space of ; that implies various restrictions on the homotopy type of.
Some of these restrictions hold in the broader context of H-spaces.
For example, the fundamental group of a topological group is abelian.

Also, for any field k, the cohomology ring has the structure of a Hopf algebra.
In view of structure theorems on Hopf algebras by Heinz Hopf and Armand Borel, this puts strong restrictions on the possible cohomology rings of topological groups.
In particular, if is a path-connected topological group whose rational cohomology ring is finite-dimensional in each degree, then this ring must be a free graded-commutative algebra over, that is, the tensor product of a polynomial ring on generators of even degree with an exterior algebra on generators of odd degree.
In particular, for a connected Lie group, the rational cohomology ring of is an exterior algebra on generators of odd degree.
Moreover, a connected Lie group has a maximal compact subgroup K, which is unique up to conjugation, and the inclusion of K into is a homotopy equivalence.
So describing the homotopy types of Lie groups reduces to the case of compact Lie groups.
For example, the maximal compact subgroup of is the circle group, and the homogeneous space can be identified with the hyperbolic plane.
Since the hyperbolic plane is contractible, the inclusion of the circle group into is a homotopy equivalence.
Finally, compact connected Lie groups have been classified by Wilhelm Killing, Élie Cartan, and Hermann Weyl.
As a result, there is an essentially complete description of the possible homotopy types of Lie groups.
For example, a compact connected Lie group of dimension at most 3 is either a torus, the group SU, or its quotient group .

Generalizations

Various generalizations of topological groups can be obtained by weakening the continuity conditions: