Group cohomology

This article is about homology and cohomology of a group. For homology or cohomology groups of a space or other object, see Homology (mathematics).

In mathematics, group cohomology is a set of mathematical tools used to study groups using cohomology theory, a technique from algebraic topology. Analogous to group representations, group cohomology looks at the group actions of a group G in an associated G-module M to elucidate the properties of the group. By treating the G-module as a kind of topological space with elements of representing n-simplices, topological properties of the space may be computed, such as the set of cohomology groups . The cohomology groups in turn provide insight into the structure of the group G and G-module M themselves. Group cohomology plays a role in the investigation of fixed points of a group action in a module or space and the quotient module or space with respect to a group action. Group cohomology is used in the fields of abstract algebra, homological algebra, algebraic topology and algebraic number theory, as well as in applications to group theory proper. As in algebraic topology, there is a dual theory called group homology. The techniques of group cohomology can also be extended to the case that instead of a G-module, G acts on a nonabelian G-group; in effect, a generalization of a module to non-Abelian coefficients.

These algebraic ideas are closely related to topological ideas. The group cohomology of a discrete group G is the singular cohomology of a suitable space having G as its fundamental group, namely the corresponding Eilenberg–MacLane space. Thus, the group cohomology of Z can be thought of as the singular cohomology of the circle S1, and similarly for Z/2Z and P(R).

A great deal is known about the cohomology of groups, including interpretations of low-dimensional cohomology, functoriality, and how to change groups. The subject of group cohomology began in the 1920s, matured in the late 1940s, and continues as an area of active research today.

Motivation

A general paradigm in group theory is that a group G should be studied via its group representations. A slight generalization of those representations are the G-modules: a G-module is an abelian group M together with a group action of G on M, with every element of G acting as an automorphism of M. We will write G multiplicatively and M additively.

Given such a G-module M, it is natural to consider the submodule of G-invariant elements:

Now, if N is a G-submodule of M (i.e. a subgroup of M mapped to itself by the action of G), it isn't in general true that the invariants in M/N are found as the quotient of the invariants in M by those in N: being invariant 'modulo N ' is broader. The purpose of the first group cohomology H1(G,N) is to precisely measure this difference.

The group cohomology functors H* in general measure the extent to which taking invariants doesn't respect exact sequences. This is expressed by a long exact sequence.

Definitions

The collection of all G-modules is a category (the morphisms are group homomorphisms f with the property f(gx) = g(f(x)) for all g in G and x in M). Sending each module M to the group of invariants MG yields a functor from the category of G-modules to the category Ab of abelian groups. This functor is left exact but not necessarily right exact. We may therefore form its right derived functors.[1] Their values are abelian groups and they are denoted by Hn(G, M), "the n-th cohomology group of G with coefficients in M". H0(G, M) is identified with MG.

Cochain complexes

The definition using derived functors is conceptually very clear, but for concrete applications, the following computations, which some authors also use as a definition, are often helpful.[2] For n ≥ 0, let Cn(G, M) be the group of all functions from Gn to M. This is an abelian group; its elements are called the (inhomogeneous) n-cochains. The coboundary homomorphisms

are defined as

One may check that , so this defines a cochain complex whose cohomology can be computed. It can be shown that the above-mentioned definition of group cohomology in terms of derived functors is isomorphic to the cohomology of this complex

Here the groups of n-cocycles, and n-coboundaries, respectively, are defined as

The functors Extn and formal definition of group cohomology

Interpreting G-modules as modules over the group ring Z[G], one can note that

i.e., the subgroup of G-invariant elements in M is identified with the group of homomorphisms from Z, which is treated as the trivial G-module (every element of G acts as the identity) to M.

Therefore, as Ext functors are the derived functors of Hom, there is a natural isomorphism

These Ext groups can also be computed via a projective resolution of Z, the advantage being that such a resolution only depends on G and not on M. We recall the definition of Ext more explicitly for this context. Let F be a projective Z[G]-resolution (e.g. a free Z[G]-resolution) of the trivial Z[G]-module Z:

e.g., one may always take the resolution of group rings, Fn = Z[Gn+1], with morphisms

Recall that for Z[G]-modules N and M, HomG(N, M) is an abelian group consisting of Z[G]-homomorphisms from N to M. Since HomG(–, M) is a contravariant functor and reverses the arrows, applying HomG(–, M) to F termwise and dropping HomG(Z,M) produces a cochain complex HomG(F, M):

The cohomology groups H*(G, M) of G with coefficients in the module M are defined as the cohomology of the above cochain complex:

for n ≥ 0.

This construction initially leads to a coboundary operator that acts on the "homogeneous" cochains. These are the elements of HomG(F, M) i.e. functions φn: GnM that obey

The coboundary operator δ: CnCn+1 is now naturally defined by, for example,

The relation to the coboundary operator d that was defined in the previous section, and which acts on the "inhomogeneous" cochains , is given by reparameterizing so that

and so on. Thus

as in the preceding section.

Group homology

Dually to the construction of group cohomology there is the following definition of group homology: given a G-module M, set DM to be the submodule generated by elements of the form g·m  m, g  G, m  M. Assigning to M its so-called coinvariants, the quotient

is a right exact functor. Its left derived functors are by definition the group homology

.

The covariant functor which assigns MG to M is isomorphic to the functor which sends M to ZZ[G] M, where Z is endowed with the trivial G-action.[3] Hence one also gets an expression for group homology in terms of the Tor functors,

Note that the superscript/subscript convention for cohomology/homology agrees with the convention for group invariants/coinvariants, while which is denoted "co-" switches:

Specifically, the homology groups Hn(G, M) can be computed as follows. Start with a projective resolution F of the trivial Z[G]-module Z, as in the previous section. Apply the covariant functor ⋅ ⊗Z[G] M to F termwise to get a chain complex FZ[G] M:

Then Hn(G, M) are the homology groups of this chain complex, for n ≥ 0.

Group homology and cohomology can be treated uniformly for some groups, especially finite groups, in terms of complete resolutions and the Tate cohomology groups.

The group homology of abelian groups G with values in a principal ideal domain k is closely related to the exterior algebra .[4]

Low-dimensional cohomology groups

H1

The first cohomology group is the quotient of the so-called crossed homomorphisms, i.e. maps (of sets) f : GM satisfying f(ab) = f(a) + af(b) for all a, b in G, modulo the so-called principal crossed homomorphisms, i.e. maps f : GM given by f(a) = amm for some fixed mM. This follows from the definition of cochains above.

If the action of G on M is trivial, then the above boils down to H1(G,M) = Hom(G, M), the group of group homomorphisms GM.

Consider the case of H1(Z/2Z,Z-), where Z- denotes the non-trivial Z/2Z-structure on the group of integers. Then crossed homomorphisms constitute all maps f: Z/2Z Z satisfying f(1)=0 and f(-1)=a for some integer a. Principal crossed homomorphisms satisfy additionally f(-1)=2a, hence

H2

If M is a trivial G-module (i.e. the action of G on M is trivial), the second cohomology group H2(G,M) is in one-to-one correspondence with the set of central extensions of G by M (up to a natural equivalence relation). More generally, if the action of G on M is nontrivial, H2(G,M) classifies the isomorphism classes of all extensions of G by M, in which the action of G on E (by inner automorphisms), endows (the image of) M with isomorphic G-module structure.

In the example as above, H2(Z/2Z,Z-)=0, as the only extensions of Z/2Z by Z are E=Z and E=Z+Z/2Z, and the action of Z/2Z (by conjugation) is trivial in both cases.

An example of a second group cohomology group is the Brauer group: it is the cohomology of the absolute Galois group of a field k which acts on the invertible elements in a separable closure:

Properties

In the following, let M be a G-module.

Long exact sequence of cohomology

In practice, one often computes the cohomology groups using the following fact: if

is a short exact sequence of G-modules, then a long exact sequence

is induced. The so-called connecting homomorphisms δn : Hn(G, N) → Hn+1(G, L) can be described in terms of inhomogeneous cochains as follows.[5] If c is an element of Hn(G, N) represented by an n-cocycle φ : Gn → N, then δn(c) is represented by dn(ψ), where ψ is an n-cochain Gn → M "lifting" φ (i.e. such that φ is the composition of ψ with the surjective map MN).

Functoriality

Group cohomology depends contravariantly on the group G, in the following sense: if f : HG is a group homomorphism, then we have a naturally induced morphism Hn(G,M) → Hn(H,M) (where in the latter, M is treated as an H-module via f). This map is called the restriction map. If the index of H in G is finite, there is also a map in the opposite direction, called transfer map,[6]

In degree 0, it is given by the map .

Given a morphism of G-modules MN, one gets a morphism of cohomology groups in the Hn(G,M) → Hn(G,N).

Products

Similarly to other cohomology theories in topology and geometry, such as singular cohomology or de Rham cohomology, group cohomology enjoys a product structure: there is a natural map called cup product:

for any two G-modules M and N. This yields a graded anti-commutative ring structure on , where R is a ring such as Z or Z/p. For a finite group G, the even part of this cohomology ring in characteristic p, carries a lot of information about the group the structure of G, for example the Krull dimension of this ring equals the maximal rank of an abelian subgroup .[7]

For example, let G be the group with two elements, under the discrete topology. The real projective space P(R) is a classifying space for G. Let k = F2, the field of two elements. Then

a polynomial k-algebra on a single generator, since this is the cellular cohomology ring of P(R).

Künneth formula

If, M = k is a field, then H*(G; k) is a graded k-algebra and the cohomology of a product of groups is related to the ones of the individual groups by a Künneth formula:

For example, if G is an elementary abelian 2-group of rank r, and k = F2, then the Künneth formula shows that the cohomology of G is a polynomial k-algebra generated by r classes in H1(G; k).,

Homology vs. cohomology

As for other cohomology theories, such as singular cohomology, group cohomology and homology are related to one another by means of a short exact sequence[8]

where A is endowed with the trivial G-action and the term at the left is the first Ext group.

Amalgamated products

Given a group A which is the subgroup of two groups G1 and G2, the homology of the amalgamated product (with coefficients in Z) lies in a long exact sequence

This can be used to compute the homology of :

is Z for n=0, Z/12 in odd degrees, and 0 otheriwse.

This exact sequence can also be applied to show that the homology of the and the special linear group agree for an infinite field k.[9]

Change of group

The Hochschild–Serre spectral sequence relates the cohomology of a normal subgroup N of G and the quotient G/N to the cohomology of the group G (for (pro-)finite groups G). From it, one gets the inflation-restriction exact sequence.

Cohomology of the classifying space

Group cohomology is closely related to topological cohomology theories such as sheaf cohomology, by means of an isomorphism

The expression BG at the left is a classifying space for G. It is an Eilenberg-MacLane space K(G,1), i.e., a space whose fundamental group is G and whose higher homotopy groups vanish).[10] Classifying spaces for Z, Z / 2 and Z / n are the 1-sphere S1, infinite real projective space , and lens spaces, respectively. In general, BG can be constructed as the quotient EG / G, where EG is a contractible space on which G acts freely. However, BG does not usually have an easily amenable geometric description.

More generally, one can attach to any G-module M a local coefficient system on BG and the above isomorphism generalizes to an isomorphism[11]

Cohomology of finite groups

Higher cohomology groups are torsion

The cohomology groups Hn(G, M) of finite groups G are all torsion for all n1. Indeed, by Maschke's theorem the category of representations of a finite group is semi-simple over any field of characteristic zero (or more generally, any field whose characteristic does not divide the order of the group), hence, viewing group cohomology as a derived functor in this abelian category, one obtains that it is zero. The other argument is that over a field of characteristic zero, the group algebra of a finite group is a direct sum of matrix algebras (possibly over division algebras which are extensions of the original field), while a matrix algebra is Morita equivalent to its base field and hence has trivial cohomology.

If the order of G is invertible in a G-module M (for example, if M is a Q-vector space), the transfer map can be used to show that the higher cohomology groups Hn(G, M) = 0 for n1. A typical application of this fact is as follows: the long exact cohomology sequence of the short exact sequence (where all three groups have a trivial G-action)

yields an isomorphism

Tate cohomology

Tate cohomology groups combine both homology and cohomology of a finite group G:

where is induced by the norm map . Tate cohomology enjoys similar features, such as long exact sequences, product structures. An important application is in class field theory, see class formation.

Tate cohomology of finite cyclic groups (G = Z/n) is 2-periodic in the sense that there are isomorphisms

A necessary and sufficient criterion for a d-periodic cohomology is that the only abelian subgroups of G are cyclic.[12] For example, any semi-direct product has this property for coprime integers n and m.

Applications

Algebraic K-theory and homology of linear groups

Algebraic K-theory is closely related to group cohomology: in Quillen's +-construction of K-theory, K-theory of a ring R is defined as the homotopy groups of a space . Here is the infinite general linear group. The space has the same homology as , i.e., the group homology of GL(R). In some cases, stability results assert that the sequence of cohomology groups

becomes stationary for large enough n, hence reducing the computation of the cohomology of the infinite general linear group to the one of some . Such results have been established when R is a field[13] or for rings of integers in a number field.[14]

The phenomenon that group homology of a series of groups stabilizes is referred to as homological stability. In addition to the case just mentioned, this applies to various other groups such as symmetric groups or mapping class groups.

Extensions

Cohomology of topological groups

Given a topological group G, i.e., a group equipped with a topology such that product and inverse are continuous maps, it is natural to consider continuous G-modules, i.e., requiring that the action

is a continuous map. For such modules, one can again consider the derived functor of . A special case occurring in algebra and number theory is when G is profinite, for example the absolute Galois group of a field. The resulting cohomology is called Galois cohomology.

Non-abelian group cohomology

Using the G-invariants and the 1-cochains, one can construct the zeroth and first group cohomology for a group G with coefficients in a non-abelian group. Specifically, a G-group is a (not necessarily abelian) group A together with an action by G.

The zeroth cohomology of G with coefficients in A is defined to be the subgroup

of elements of A fixed by G.

The first cohomology of G with coefficients in A is defined as 1-cocycles modulo an equivalence relation instead of by 1-coboundaries. The condition for a map φ to be a 1-cocycle is that φ(gh) = φ(g)[gφ(h)] and if there is an a in A such that . In general, H1(G, A) is not a group when A is non-abelian. It instead has the structure of a pointed set – exactly the same situation arises in the 0th homotopy group, which for a general topological space is not a group but a pointed set. Note that a group is in particular a pointed set, with the identity element as distinguished point.

Using explicit calculations, one still obtains a truncated long exact sequence in cohomology. Specifically, let

be a short exact sequence of G-groups, then there is an exact sequence of pointed sets

History and relation to other fields

The low-dimensional cohomology of a group was classically studied in other guises, long before the notion of group cohomology was formulated in 1943–45. The first theorem of the subject can be identified as Hilbert's Theorem 90 in 1897; this was recast into Noether's equations in Galois theory (an appearance of cocycles for H1). The idea of factor sets for the extension problem for groups (connected with H2) arose in the work of Hölder (1893), in Issai Schur's 1904 study of projective representations, in Schreier's 1926 treatment, and in Richard Brauer's 1928 study of simple algebras and the Brauer group. A fuller discussion of this history may be found in (Weibel 1999, pp. 806–811).

In 1941, while studying H2(G, Z) (which plays a special role in groups), Hopf discovered what is now called Hopf's integral homology formula (Hopf 1942), which is identical to Schur's formula for the Schur multiplier of a finite, finitely presented group:

,

where GF/R and F is a free group.

Hopf's result led to the independent discovery of group cohomology by several groups in 1943-45: Eilenberg and Mac Lane in the USA (Rotman 1995, p. 358); Hopf and Eckmann in Switzerland; and Freudenthal in the Netherlands (Weibel 1999, p. 807). The situation was chaotic because communication between these countries was difficult during World War II.

From a topological point of view, the homology and cohomology of G was first defined as the homology and cohomology of a model for the topological classifying space BG as discussed above. In practice, this meant using topology to produce the chain complexes used in formal algebraic definitions. From a module-theoretic point of view this was integrated into the CartanEilenberg theory of homological algebra in the early 1950s.

The application in algebraic number theory to class field theory provided theorems valid for general Galois extensions (not just abelian extensions). The cohomological part of class field theory was axiomatized as the theory of class formations. In turn, this led to the notion of Galois cohomology and étale cohomology (which builds on it) (Weibel 1999, p. 822). Some refinements in the theory post-1960 have been made, such as continuous cocycles and Tate's redefinition, but the basic outlines remain the same. This is a large field, and now basic in the theories of algebraic groups.

The analogous theory for Lie algebras, called Lie algebra cohomology, was first developed in the late 1940s, by Chevalley–Eilenberg, and Koszul (Weibel 1999, p. 810). It is formally similar, using the corresponding definition of invariant for the action of a Lie algebra. It is much applied in representation theory, and is closely connected with the BRST quantization of theoretical physics.

Group cohomology theory also has a direct application in condensed matter physics. Just like group theory being the mathematical foundation of spontaneous symmetry breaking phases, group cohomology theory is the mathematical foundation of a class of quantum states of matter—short-range entangled states with symmetry. Short-range entangled states with symmetry are also known as symmetry protected topological states.

Notes

  1. This uses that the category of G-modules has enough injectives, since it is isomorphic to the category of all modules over the group ring Z[G].
  2. Page 62 of Milne 2008 or section VII.3 of Serre 1979
  3. Recall that the tensor product NZ[G] M is defined whenever N is a right Z[G]-module and M is a left Z[G]-module. If N is a left Z[G]-module, we turn it into a right Z[G]-module by setting a g = g−1 a for every gG and every aN. This convention allows to define the tensor product NZ[G] M in the case where both M and N are left Z[G]-modules.
  4. For example, the two are isomorphic if all primes p such that G has p-torsion are invertible in k. See (Knudson 2001), Theorem A.1.19 for the precise statement.
  5. Remark II.1.21 of Milne 2008
  6. (Brown 1972), §III.9
  7. Quillen, Daniel. The spectrum of an equivariant cohomology ring. I. II. Ann. Math. (2) 94, 549-572, 573-602 (1971).
  8. (Brown 1972), Exercise III.1.3
  9. (Knudson 2001), Chapter 4
  10. For this, G is assumed to be discrete. For general topological groups, .
  11. (Adem & Milgram 2004), Chapter II.
  12. (Brown 1972), §VI.9
  13. A. A. Suslin (1984), "Homology of GLn, characteristic classes and Milnor K-theory.", AlgebraicK-theory, number theory, geometry and analysis, Lecture Notes in Math., 1046, Springer, pp. 357–375
  14. In this case, the coefficients are rational. Borel, Armand (1974), "Stable real cohomology of arithmetic groups", Annales scientifiques de l'École Normale Supérieure, Series 4, 7 (2): 235–272

References

This article is issued from Wikipedia - version of the 11/9/2016. The text is available under the Creative Commons Attribution/Share Alike but additional terms may apply for the media files.