Group cohomology

Last updated

In mathematics (more specifically, in homological algebra), 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.

Contents

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 can be thought of as the singular cohomology of the circle S1, and similarly for and

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 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 is to precisely measure this difference.

The group cohomology functors 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 for all g in G and x in M). Sending each module M to the group of invariants 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. [lower-alpha 1] Their values are abelian groups and they are denoted by , "the n-th cohomology group of G with coefficients in M". Furthermore, the group may be identified with .

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. [1] For let be the group of all functions from to M (here means ). This is an abelian group; its elements are called the (inhomogeneous) n-cochains. The coboundary homomorphisms are defined by

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 one can note that

i.e., the subgroup of G-invariant elements in M is identified with the group of homomorphisms from , 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 , 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 -resolution (e.g. a free -resolution) of the trivial -module :

e.g., one may always take the resolution of group rings, with morphisms

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

The cohomology groups of G with coefficients in the module M are defined as the cohomology of the above cochain complex:

This construction initially leads to a coboundary operator that acts on the "homogeneous" cochains. These are the elements of , that is, functions that obey

The coboundary operator 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 where is endowed with the trivial G-action. [lower-alpha 2] 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 -module as in the previous section. Apply the covariant functor to F termwise to get a chain complex :

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 . [lower-alpha 3]

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(g) = gmm 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, since the crossed homomorphisms are then just ordinary homomorphisms and the coboundaries (i.e. the principal crossed homomorphisms) must have image identically zero: hence there is only the zero coboundary.

On the other hand, consider the case of where denotes the non-trivial-structure on the additive group of integers, which sends a to -a for every ; and where we regard as the group . By considering all possible cases for the images of , it may be seen that crossed homomorphisms constitute all maps satisfying and for some arbitrary choice of integer t. Principal crossed homomorphisms must additionally satisfy for some integer m: hence every crossed homomorphism sending -1 to an even integer is principal, and therefore:

with the group operation being pointwise addition: , noting that is the identity element.

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 an isomorphic G-module structure.

In the example from the section on immediately above, as the only extension of by with the given nontrivial action is the infinite dihedral group, which is a split extension and so trivial inside the group. This is in fact the significance in group-theoretical terms of the unique non-trivial element of .

An example of a second 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:

See also .

Basic examples

Group cohomology of a finite cyclic group

For the finite cyclic group of order with generator , the element in the associated group ring is a divisor of zero because its product with , given by

gives

This property can be used to construct the resolution [2] [3] of the trivial -module via the complex

giving the group cohomology computation for any -module . Note the augmentation map gives the trivial module its -structure by

This resolution gives a computation of the group cohomology since there is the isomorphism of cohomology groups

showing that applying the functor to the complex above (with removed since this resolution is a quasi-isomorphism), gives the computation

for

For example, if , the trivial module, then , , and , hence

Explicit cocycles

Cocycles for the group cohomology of a cyclic group can be given explicitly [4] prop 2.3 using the Bar resolution. We get a complete set of generators of -cocycles for odd as the maps

given by

for odd, , a primitive -th root of unity, a field containing -th roots of unity, and

for a rational number denoting the largest integer not greater than . Also, we are using the notation

where is a generator for . Note that for non-zero even indices the cohomology groups are trivial.

Cohomology of free groups

Using a resolution

Given a set the associated free group has an explicit resolution [5] of the trivial module which can be easily computed. Notice the augmentation map

has kernel given by the free submodule generated by the set , so

.

Because this object is free, this gives a resolution

hence the group cohomology of with coefficients in can be computed by applying the functor to the complex , giving

this is because the dual map

sends any -module morphism

to the induced morphism on by composing the inclusion. The only maps which are sent to are -multiples of the augmentation map, giving the first cohomology group. The second can be found by noticing the only other maps

can be generated by the -basis of maps sending for a fixed , and sending for any .

Using topology

The group cohomology of free groups generated by letters can be readily computed by comparing group cohomology with its interpretation in topology. Recall that for every group there is a topological space , called the classifying space of the group, which has the property

In addition, it has the property that its topological cohomology is isomorphic to group cohomology

giving a way to compute some group cohomology groups. Note could be replaced by any local system which is determined by a map

for some abelian group . In the case of for letters, this is represented by a wedge sum of circles [6] which can be showed using the Van-Kampen theorem, giving the group cohomology [7]

Group cohomology of an integral lattice

For an integral lattice of rank (hence isomorphic to ), its group cohomology can be computed with relative ease. First, because , and has , which as abelian groups are isomorphic to , the group cohomology has the isomorphism

with the integral cohomology of a torus of rank .

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,

can be described in terms of inhomogeneous cochains as follows. [8] If is represented by an n-cocycle then is represented by where is an n-cochain "lifting" (i.e. is the composition of 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, [9]

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 or 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 . [10]

For example, let G be the group with two elements, under the discrete topology. The real projective space is a classifying space for G. Let the field of two elements. Then

a polynomial k-algebra on a single generator, since this is the cellular cohomology ring of

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 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 [11]

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 integer coefficients) lies in a long exact sequence

The homology of can be computed using this:

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. [12]

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 [13]

The expression at the left is a classifying space for . It is an Eilenberg–MacLane space , i.e., a space whose fundamental group is and whose higher homotopy groups vanish). [lower-alpha 4] Classifying spaces for and are the 1-sphere S1, infinite real projective space and lens spaces, respectively. In general, can be constructed as the quotient , where is a contractible space on which acts freely. However, does not usually have an easily amenable geometric description.

More generally, one can attach to any -module a local coefficient system on and the above isomorphism generalizes to an isomorphism [14]

Further examples

Semi-direct products of groups

There is a way to compute the semi-direct product of groups using the topology of fibrations and properties of Eilenberg-Maclane spaces. Recall that for a semi-direct product of groups there is an associated short exact sequence of groups

Using the associated Eilenberg-Maclane spaces there is a Serre fibration

which can be put through a Serre spectral sequence. This gives an -page

which gives information about the group cohomology of from the group cohomology groups of . Note this formalism can be applied in a purely group-theoretic manner using the Lyndon–Hochschild–Serre spectral sequence.

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 -vector space), the transfer map can be used to show that for 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, 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. [15] 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 [16] or for rings of integers in a number field. [17]

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.

Projective representations and group extensions

In quantum mechanics we often have systems with a symmetry group We expect an action of on the Hilbert space by unitary matrices We might expect but the rules of quantum mechanics only require

where is a phase. This projective representation of can also be thought of as a conventional representation of a group extension of by as described by the exact sequence

Requiring associativity

leads to

which we recognise as the statement that i.e. that is a cocycle taking values in We can ask whether we can eliminate the phases by redefining

which changes

This we recognise as shifting by a coboundary The distinct projective representations are therefore classified by Note that if we allow the phases themselves to be acted on by the group (for example, time reversal would complex-conjugate the phase), then the first term in each of the coboundary operations will have a acting on it as in the general definitions of coboundary in the previous sections. For example,

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 to be a 1-cocycle is that and if there is an a in A such that . In general, 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, well 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 Emmy Noether's equations in Galois theory (an appearance of cocycles for ). The idea of factor sets for the extension problem for groups (connected with ) arose in the work of Otto Hölder (1893), in Issai Schur's 1904 study of projective representations, in Otto 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 (which plays a special role in groups), Heinz 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 and F is a free group.

Hopf's result led to the independent discovery of group cohomology by several groups in 1943-45: Samuel Eilenberg and Saunders Mac Lane in the United States ( Rotman 1995 , p. 358); Hopf and Beno Eckmann in Switzerland; Hans Freudenthal in the Netherlands ( Weibel 1999 , p. 807); and Dmitry Faddeev in the Soviet Union (Arslanov 2011 , p. 29, Faddeev 1947). 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 John 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 Claude Chevalley and Eilenberg, and Jean-Louis 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. [18] [19]

See also

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
  2. Recall that the tensor product is defined whenever N is a right -module and M is a left -module. If N is a left -module, we turn it into a right -module by setting ag = g−1a for every gG and every aN. This convention allows to define the tensor product in the case where both M and N are left -modules.
  3. 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.
  4. For this, G is assumed to be discrete. For general topological groups, .

Related Research Articles

<span class="mw-page-title-main">Associative algebra</span> Algebraic structure with (a + b)(c + d) = ac + ad + bc + bd and (a)(bc) = (ab)(c)

In mathematics, an associative algebraA is an algebraic structure with compatible operations of addition, multiplication, and a scalar multiplication by elements in some field K. The addition and multiplication operations together give A the structure of a ring; the addition and scalar multiplication operations together give A the structure of a vector space over K. In this article we will also use the term K-algebra to mean an associative algebra over the field K. A standard first example of a K-algebra is a ring of square matrices over a field K, with the usual matrix multiplication.

<span class="mw-page-title-main">Homological algebra</span> Branch of mathematics

Homological algebra is the branch of mathematics that studies homology in a general algebraic setting. It is a relatively young discipline, whose origins can be traced to investigations in combinatorial topology and abstract algebra at the end of the 19th century, chiefly by Henri Poincaré and David Hilbert.

In mathematics, homology is a general way of associating a sequence of algebraic objects, such as abelian groups or modules, with other mathematical objects such as topological spaces. Homology groups were originally defined in algebraic topology. Similar constructions are available in a wide variety of other contexts, such as abstract algebra, groups, Lie algebras, Galois theory, and algebraic geometry.

In mathematics, specifically in homology theory and algebraic topology, cohomology is a general term for a sequence of abelian groups, usually one associated with a topological space, often defined from a cochain complex. Cohomology can be viewed as a method of assigning richer algebraic invariants to a space than homology. Some versions of cohomology arise by dualizing the construction of homology. In other words, cochains are functions on the group of chains in homology theory.

In mathematics, a sheaf is a tool for systematically tracking data attached to the open sets of a topological space and defined locally with regard to them. For example, for each open set, the data could be the ring of continuous functions defined on that open set. Such data is well behaved in that it can be restricted to smaller open sets, and also the data assigned to an open set is equivalent to all collections of compatible data assigned to collections of smaller open sets covering the original open set.

In algebraic geometry, motives is a theory proposed by Alexander Grothendieck in the 1960s to unify the vast array of similarly behaved cohomology theories such as singular cohomology, de Rham cohomology, etale cohomology, and crystalline cohomology. Philosophically, a "motif" is the "cohomology essence" of a variety.

In algebraic topology, singular homology refers to the study of a certain set of algebraic invariants of a topological space X, the so-called homology groups Intuitively, singular homology counts, for each dimension n, the n-dimensional holes of a space. Singular homology is a particular example of a homology theory, which has now grown to be a rather broad collection of theories. Of the various theories, it is perhaps one of the simpler ones to understand, being built on fairly concrete constructions.

In mathematics, certain functors may be derived to obtain other functors closely related to the original ones. This operation, while fairly abstract, unifies a number of constructions throughout mathematics.

In mathematics, Kähler differentials provide an adaptation of differential forms to arbitrary commutative rings or schemes. The notion was introduced by Erich Kähler in the 1930s. It was adopted as standard in commutative algebra and algebraic geometry somewhat later, once the need was felt to adapt methods from calculus and geometry over the complex numbers to contexts where such methods are not available.

In mathematics, the Ext functors are the derived functors of the Hom functor. Along with the Tor functor, Ext is one of the core concepts of homological algebra, in which ideas from algebraic topology are used to define invariants of algebraic structures. The cohomology of groups, Lie algebras, and associative algebras can all be defined in terms of Ext. The name comes from the fact that the first Ext group Ext1 classifies extensions of one module by another.

In mathematics, particularly category theory, a representable functor is a certain functor from an arbitrary category into the category of sets. Such functors give representations of an abstract category in terms of known structures allowing one to utilize, as much as possible, knowledge about the category of sets in other settings.

In mathematics, specifically algebraic topology, an Eilenberg–MacLane space is a topological space with a single nontrivial homotopy group.

In mathematics, Lie algebra cohomology is a cohomology theory for Lie algebras. It was first introduced in 1929 by Élie Cartan to study the topology of Lie groups and homogeneous spaces by relating cohomological methods of Georges de Rham to properties of the Lie algebra. It was later extended by Claude Chevalley and Samuel Eilenberg (1948) to coefficients in an arbitrary Lie module.

In mathematics, the tensor product of modules is a construction that allows arguments about bilinear maps to be carried out in terms of linear maps. The module construction is analogous to the construction of the tensor product of vector spaces, but can be carried out for a pair of modules over a commutative ring resulting in a third module, and also for a pair of a right-module and a left-module over any ring, with result an abelian group. Tensor products are important in areas of abstract algebra, homological algebra, algebraic topology, algebraic geometry, operator algebras and noncommutative geometry. The universal property of the tensor product of vector spaces extends to more general situations in abstract algebra. The tensor product of an algebra and a module can be used for extension of scalars. For a commutative ring, the tensor product of modules can be iterated to form the tensor algebra of a module, allowing one to define multiplication in the module in a universal way.

In algebraic geometry, local cohomology is an algebraic analogue of relative cohomology. Alexander Grothendieck introduced it in seminars in Harvard in 1961 written up by Hartshorne (1967), and in 1961-2 at IHES written up as SGA2 - Grothendieck (1968), republished as Grothendieck (2005). Given a function defined on an open subset of an algebraic variety, local cohomology measures the obstruction to extending that function to a larger domain. The rational function , for example, is defined only on the complement of on the affine line over a field , and cannot be extended to a function on the entire space. The local cohomology module detects this in the nonvanishing of a cohomology class . In a similar manner, is defined away from the and axes in the affine plane, but cannot be extended to either the complement of the -axis or the complement of the -axis alone ; this obstruction corresponds precisely to a nonzero class in the local cohomology module .

In mathematics, the Adams spectral sequence is a spectral sequence introduced by J. Frank Adams (1958) which computes the stable homotopy groups of topological spaces. Like all spectral sequences, it is a computational tool; it relates homology theory to what is now called stable homotopy theory. It is a reformulation using homological algebra, and an extension, of a technique called 'killing homotopy groups' applied by the French school of Henri Cartan and Jean-Pierre Serre.

In mathematics, a Hodge structure, named after W. V. D. Hodge, is an algebraic structure at the level of linear algebra, similar to the one that Hodge theory gives to the cohomology groups of a smooth and compact Kähler manifold. Hodge structures have been generalized for all complex varieties in the form of mixed Hodge structures, defined by Pierre Deligne (1970). A variation of Hodge structure is a family of Hodge structures parameterized by a manifold, first studied by Phillip Griffiths (1968). All these concepts were further generalized to mixed Hodge modules over complex varieties by Morihiko Saito (1989).

In mathematics, Hochschild homology (and cohomology) is a homology theory for associative algebras over rings. There is also a theory for Hochschild homology of certain functors. Hochschild cohomology was introduced by Gerhard Hochschild (1945) for algebras over a field, and extended to algebras over more general rings by Henri Cartan and Samuel Eilenberg (1956).

In algebraic geometry, a presheaf with transfers is, roughly, a presheaf that, like cohomology theory, comes with pushforwards, “transfer” maps. Precisely, it is, by definition, a contravariant additive functor from the category of finite correspondences to the category of abelian groups.

In mathematics, and especially algebraic geometry, a Bridgeland stability condition, defined by Tom Bridgeland, is an algebro-geometric stability condition defined on elements of a triangulated category. The case of original interest and particular importance is when this triangulated category is the derived category of coherent sheaves on a Calabi–Yau manifold, and this situation has fundamental links to string theory and the study of D-branes.

References

  1. Page 62 of Milne 2008 or section VII.3 of Serre 1979
  2. Dummit, David Steven; Foote, Richard M. (14 July 2003). Abstract algebra (Third ed.). Hoboken, NJ. p. 801. ISBN   0-471-43334-9. OCLC   52559229.{{cite book}}: CS1 maint: location missing publisher (link)
  3. Brown, Kenneth S. (6 December 2012). Cohomology of groups. New York, New York. p. 35. ISBN   978-1-4684-9327-6. OCLC   853269200.{{cite book}}: CS1 maint: location missing publisher (link)
  4. Huang, Hua-Lin; Liu, Gongxiang; Ye, Yu (2012-06-23). "The braided monoidal structures on a class of linear Gr-categories". arXiv: 1206.5402v3 .{{cite journal}}: Cite journal requires |journal= (help)
  5. Evens, Leonard. (1991). The cohomology of groups. Oxford: Clarendon Press. ISBN   0-19-853580-5. OCLC   23732584.
  6. Hatcher, Allen (2002). Algebraic topology. Cambridge: Cambridge University Press. p. 43. ISBN   0-521-79160-X. OCLC   45420394.
  7. Webb, Peter. "An Introduction to the Cohomology of Groups" (PDF). Archived (PDF) from the original on 6 May 2020.
  8. Remark II.1.21 of Milne 2008
  9. ( Brown 1972 ), §III.9
  10. Quillen, Daniel. The spectrum of an equivariant cohomology ring. I. II. Ann. Math. (2) 94, 549-572, 573-602 (1971).
  11. ( Brown 1972 ), Exercise III.1.3
  12. ( Knudson 2001 ), Chapter 4
  13. Stasheff, James D. (1978-07-01). "Continuous cohomology of groups and classifying spaces". Bulletin of the American Mathematical Society. 84 (4): 513–531. doi: 10.1090/s0002-9904-1978-14488-7 . ISSN   0002-9904.
  14. ( Adem & Milgram 2004 ), Chapter II.
  15. ( Brown 1972 ), §VI.9
  16. Suslin, Andrei A. (1984), "Homology of , characteristic classes and Milnor K-theory", Algebraic K-theory, number theory, geometry and analysis, Lecture Notes in Mathematics, vol. 1046, Springer, pp. 357–375
  17. In this case, the coefficients are rational. Borel, Armand (1974). "Stable real cohomology of arithmetic groups". Annales Scientifiques de l'École Normale Supérieure . Série 4. 7 (2): 235–272. doi: 10.24033/asens.1269 .
  18. Wang, Juven C.; Gu, Zheng-Cheng; Wen, Xiao-Gang (22 January 2015). "Field-Theory Representation of Gauge-Gravity Symmetry-Protected Topological Invariants, Group Cohomology, and Beyond". Physical Review Letters. 114 (3): 031601. arXiv: 1405.7689 . Bibcode:2015PhRvL.114c1601W. doi:10.1103/physrevlett.114.031601. ISSN   0031-9007. PMID   25658993. S2CID   2370407.
  19. Wen, Xiao-Gang (4 May 2015). "Construction of bosonic symmetry-protected-trivial states and their topological invariants via G×SO(∞) nonlinear σ models". Physical Review B. 91 (20): 205101. arXiv: 1410.8477 . Bibcode:2015PhRvB..91t5101W. doi:10.1103/physrevb.91.205101. ISSN   1098-0121. S2CID   13950401.

Works cited

Further reading