<<

AKHIL MATHEW

1. COHOMOLOGYAND

Let G be a group. We can form the Z[G] over G; by definition it is the set P of formal finite sums aigi, where ai ∈ Z, gi ∈ G, and multiplication is defined in the obvious manner. We shall call an A a G- if it is a left Z[G]-module. This means, of course, that there exists a homomorphism G → AutZ(A). We can also make A into a right Z[G]-module simply by writing ag := g−1a for a ∈ A, g ∈ G. This is important for tensor products. An example of a G-module is any abelian group with trivial action by G. For instance, we shall in the future denote by Z the integers with trivial G-action. Finally, if A and B are G-modules, then a G-homomorphism between them is a map φ : A → B which is a Z[G] homomorphism. The set of G-homomorphisms between A and B is denoted by HomG(A, B). It is a left exact of A and B, covariant in B and contravariant in A. As usual its derived are denoted by Exti. Let A be a G-module. Then we define the cohomology groups as Hi(G, A) := Exti ( ,A). Z[G] Z Then the Hi(G, ·) are covariant functors from the category of G-modules to the category 0 of abelian groups. Now we have clearly H (G, A) = HomG(Z,A), by basic properties of Ext over any ring. Also, a Z-homomorphism Z → A is determined by its value at 1; it is a G-homomorphism if and only if its image a ∈ A is fixed by G, i.e ga = a for all G G g ∈ G. Denote the set of such a by A . So we see that HomG(Z,A) = A ; in particular A → AG is a left . Then another way of stating our definitions is that Hi(G, ·) are the derived functors of the functor A → AG. i Go back to the initial definition of H (G, ·) in terms of Ext. Let {Pi} be a projective of Z, i.e. an · · · → Pn → Pn−1 → · · · → P1 → P0 → Z, where all the Pi are projective (e.g. free) G-modules. Then we have a

0 → HomG(P0,A) → HomG(P1,A) .... The homology groups of this complex are the derived functors Hi(G, A). The above characterization yields the following:

Proposition 1. Let A be a co-induced G-module, i.e. A = HomZ(Z[G],B) for a Z- module B. (This can be made into a G-module by multiplication on the right.) Then we have Hi(G, A) = 0 if i > 0.

Proof. Consider the chain complex HomG(Pi,A) = HomG(Pi, HomZ(Z[G],B)) = HomZ(Pi,B), as is easily seen. Since the Pi are projective, this sequence is exact except possibly at i = 0, and the proposition follows. 

Date: January 25, 2009. 1 2 AKHIL MATHEW

Moreover, by basic properties of derived functors, we have the following result: If 0 → A → B → C → 0 is a “short” exact sequence of G-modules, then there exists a “long” exact sequence 0 → AG → BG → CG → H1(G, A) → H1(G, B) → H1(G, C) → H2(G, A) → H2(G, B) .... The definition of homology is similar. Let A be a G-module. We define

Hi(G, A) := TorZ[G](Z,A). Here we have used the making of A into a right Z[G]-module, as above. Then by defini- tion, the Hi(G, ·) are the derived functors of the functor A → Z ⊗Z[G] A, which is also H0(G, A), by elementary properties of Tor.

In the future, we suppress the Z[G] in the tensor product ⊗Z[G], for convenience. Let us compute the functor Z ⊗Z[G] A. We have an exact sequence of G-modules 0 → IG → Z[G] → Z → 0, where IG is the augmentation ideal, i.e. the G-submodule of Z[G] generated by the g − 1, g ∈ G; and the homomorphism Z[G] → Z is the augmentation homomorphism that maps each g to 1. Since the tensor product is right exact, we have the following exact sequence:

IG ⊗Z[G] A → Z[G] ⊗Z[G] A → Z ⊗Z[G] A → 0. The second term is just A. Thus we see that Z⊗A = A/IGA. In other words, the Hi(G, ·) are the derived functors of A → A/IGA. There are properties of the homology groups. As usual, if {Pi} is a projective resolution of Z as before, then the Hi(G, A) are the homology groups of the chain complex (tensor products over Z[G])

· · · → P2 ⊗ A → P1 ⊗ A → P0 ⊗ A → 0. This immediately yields the following analog of Proposition 1:

Proposition 2. Let A be an induced G-module, i.e. A = Z[G] ⊗Z B for a Z-module B. (This can be made into a G-module by multiplication on the left.) Then we have Hi(G, A) = 0 if i > 0. Proof. Indeed:

Pn ⊗Z[G] A = Pn ⊗Z[G] (Z[G] ⊗Z B) = (Pn ⊗Z[G] Z[G]) ⊗Z B = Pn ⊗Z B, and the Pn are projective.  We have a similar long exact sequence, which we leave the reader to write out.

2. EXAMPLES

We construct an explicit resolution of Z, which will give us a practical criterion for computing the cohomology and homology groups. Define Pi to be the free abelian group generated by the symbols (g0, . . . , gi) for gj ∈ G, 0 ≤ j ≤ i. Let G act on Pi in the obvious manner, i.e. g(g0, . . . , gi) := (gg0, . . . , ggi). Then the Pi are G-modules. We give the boundary homomorphisms. The homomorphism P0 → Z maps (g0) → 1 for all g0 ∈ G, and is extended by linearity. The homomorphism di : Pi → Pi−1 for i > 0 is given by i X j di(g0, . . . , gi) := (−1) (g0,..., gˆj, . . . , gi), j=0 GROUP COHOMOLOGY 3 where as usual the hat means a term is omitted. These are G-homomorphisms and form a chain complex (compute di ◦ di−1 directly). We must now check exactness. First, if d0(a) = 0, then we can write a as a sum of P (gj) − (hj), which is the boundary of (gj, hj). Next, fix g ∈ G. We define a chain homotopy Di : Pi → Pi+1 by Di(g0, . . . , gi) := (g, g0, . . . , gi). We see easily that Dd + dD = 1, by abuse of notation, whence the sequence is exact. In the future we will drop indexes like that. Hence, for instance, a n-cocycle in some G-module A is a G-homomorphism f : Pn → A such that fd = 0. f is a n-coboundary if f = gd for some G-homomorphism Pn−1 → A. There is, however, a more convenient notation. Pi is a free G-module on the set [g1, . . . , gi] := (1, g1, g1g2, g1g2g3, . . . , g1 . . . gi). So a G-homomorphism f : Pi → A is equivalent to a map assigning an element of A to each [g1, . . . , gi], which extends to the free module in the usual manner. Note that i−1 X j i d[g1, . . . , gi] = g1[g2, . . . , gi] + (−1) [g1, . . . , gjgj+1, . . . , gi] + (−1) [g1, . . . , gi−1]. j=1

Now denote the set of [g1, . . . , gi] by Si. Hence a n-cocycle becomes a map f : Si → A such that i X j i+1 g1f([g2, . . . , gi+1])+ (−1) f([g1, . . . , gjgj+1, . . . , gi])+(−1) f([g1, . . . , gi]) = 0, j=1 where a n-coboundary becomes a map f such that there exists a map h : Si−1 → A with i−1 X j i f([g1, . . . , gi]) = g1h([g2, . . . , gi])+ (−1) h([g1, . . . , gjgj+1, . . . , gi])+(−1) h([g1, . . . , gi−1]. j=1 Take the case i = 1. Then a 1-cocyle is a map f : G → A such that f(st) = sf(t) + f(s), or a crossed homomorphism. A 1-coboundary is a map f of the form f(s) = sa − a. This allows us to compute the cohomology groups in special cases. We will use one more example, this time for homology:

Proposition 3. H1(G, Z) = Gab, the abelianization of G. Proof. We have an exact sequence

0 → IG → Z[G] → Z → 0, with the usual augmentation ideal and augmentation homomorphism. Taking the long ex- act sequence and noting that Z[G] is Z[G]-free (hence projective), we see that H1(G, Z) = 2 2 H0(G, IG) = IG/IG. So we have to define a map f : Gab → IG/IG which is an isomor- phism. Let f(s) := s − 1; then f(st) − f(s) − f(t) = st − 1 − (s − 1) − (t − 1) = 0 in 2 2 IG, so f is a homomorphism. For the inverse, define g : IG/IG → Gab by g(σ − 1) := σ. 2 Note that all elements in IG/IG are of that form. As above this is a homomorphism (since Gab is abelian), and it is clearly the inverse of f. 

3. INFLATION,RESTRICTION,CORESTRICTION Let A be a G-module, and let f : G0 → G be a homomorphism of groups. We can consider A as a G0 module by setting g0a := f(g0)a for g0 ∈ G0. We have obviously 0 AG ⊂ AG . Hence there is an inclusion homomorphism H0(G, A) → H0(G0,A). Now the families of functors {Hi(G, ·)} and {Hi(G0, ·)} are universal δ-functors in the sense 4 AKHIL MATHEW of Grothendieck, (see [?]) since they are derived functors. Hence this morphism extends uniquely to morphisms Hi(G, A) → Hi(G0,A), i ≥ 0, which are, as usual, functorial in A, and which commute with the boundary homomor- phisms for long exact sequences generated by short exact sequences 0 → A → B → C → 0. If H ⊂ G is a subgroup, and A is a H module then we get a map Res : Hi(G, A) → Hi(H,A). This is called restriction. We can of course define it directly on the cochains and cobound- aries, by a pull-back. That this gives the same restriction map is immediate from the uniqueness. We can go in the opposite direction. Let H ⊂ G be a normal subgroup, and let A be a G-module. Then AH is a G/H module in the obvious manner. Indeed, if a ∈ A is fixed by the actions of H, and g ∈ G, h ∈ H, we have that h(ga) = gh0a = ga for some h0 ∈ H, since H is normal. This shows that AH is a G/H-module. The canonical homomorphism G → G/H induces a homomorphism Hi(G/H, AH ) → Hi(G, AH ), which when composed with the canonical map Hi(G, AH ) → Hi(G, A) induced by in- clusion, yields the inflation map: Hi(G/H, AH ) → Hi(G, A), which is, of course, a morphism of δ-functors as well. Both restriction and inflation can be computed directly via cochains as well, in an obvi- ous manner. Proposition 4 (Restriction-Inflation Exact Sequence). Let A, H, G be as above. Then the sequence with inflation and restriction 0 → H1(G/H, AH ) → H1(G, A) → H1(H,A) is exact. Proof. Direct verification on cochains and coboundaries. (1) We first check that the first map is injective. If f : G/H → AH is a cochain, then the pull-back fG : G → A represents Inf(f). If fG(s) = sa − a for some a ∈ A, i.e. fG is a coboundary, then it follows immediately from the fact that fG is constant on cosets of H that 1a − a = t1a − a for t ∈ H, or a ∈ AH . It thus follows that f itself is a coboundary (through a). (2) Now we check that the composite of the two maps is zero. So choose a crossed homomorphism f : G/H → AH . Since f(11)˙ = f(1) + 1f(1) = 2f(1), we see that f(1) = 0. Lifting to G and denoting the inflation by fG we see that fG(s) = 0 for s ∈ H; restricting to H we get the zero cocyle. (3) Now all we need is to show that if h : G → A is a crossed homomorpism and there exists a ∈ A such that h(s) = sa − a for s ∈ H, then h is an inflation of some crossed homomorphism f : G/H → A. By subtracting the coboundary sa − a from h, we assume h vanishes on H. Then we have h(st) = h(s) + sh(t) for any s, t ∈ G. Choose t ∈ H so that h(t) = 0; then h(st) = h(s). We see that h is constant on the H-cosets. Now take s ∈ H, and we get h(t) = h(st) = 0 + sh(t). GROUP COHOMOLOGY 5

Thus h(t) is fixed under H, so the values of H come from AH . This is enough to conclude that h is an inflation and concludes the proof.  Proposition 1. Let H ⊂ G be normal. Let q > 1, and suppose that Hi(H,A) = 0 if 1 ≤ i ≤ i − 1. Then we have an exact sequence: 0 → Hq(G/H, AH ) → Hq(G/H, A) → Hq(H,A). Proof. This follows from a simple dimension-shifting argument. Embed A in a G-coinduced module B; let C be the cokernel. Then B is also H-coinduced. Also BH is G/H coin- duced. Moreover, we have an exact sequence 0 → AH → BH → CH → 0, by looking at the H-cohomology. One uses the commutative diagram: 0 −−−−→ Hq−1(G/H, CH ) −−−−→ Hq−1(G, C) −−−−→ Hq−1(H,C)       y y y 0 −−−−→ Hq−1(G/H, AH ) −−−−→ Hq−1(G, A) −−−−→ Hq−1(A, C). The vertical boundary maps are isomorphisms, hence the proposition follows by induction.  There is a similar definition for corestriction. Let A be a G-module, H ⊂ G not necessarily normal. One begins with the obvious map

A/IH A → A/IGA, and one gets morphisms Hi(H,A) → Hi(G, A), for all i, which are natural in the usual senses. We can also define corestriction on the cohomology groups. Let H ⊂ G. Then AG ⊂ AH . Let S be a set of representatives of left cosets of H in G, so that SH = G. Define f : AH → AG by: X f(a) := sa. s∈S Then f defines the corestriction on the zeroth cohomology group. The map is clearly independent of the choice of S. It extends to the corestriction maps on all the groups. There is a similar procedure for the restriction on the homology groups. Proposition 2. We have Cor ◦ Res = (G : H) on all Hi(G, A) (multiplication by (G : H)).

Proof. Trivial for i = 0. Dimension-shifting for i > 0. 

4. THE TATE GROUPS P Let G be a finite group, and A a G-module. We define N = g∈G g ∈ Z[G]. G Clearly we have NA ⊂ A , and IGA ⊂ AN . (We use the notation Af for a ho- momorphism of A into some other group to denote ker f.) We get a homomorphism 0 0 NA : H0(G, A) → H (G, A). Using this, we can splie the homology and cohomology groups together. Indeed, we set: (1) Hˆ i(G, A) := Hi(G, A) if i > 0. 6 AKHIL MATHEW

ˆ 0 G 0 (2) H (G, A) := A /NA = CokerNA. ˆ −1 0 (3) H (G, A) := AN /IGA = KerNA. i (4) Hˆ (G, A) := Hi+1(G, A) if i < −1. These are called the Tate cohomology groups. They form a δ-functor in the following manner. Let 0 → A → B → C → 0 be a short exact sequence. We have the commutative diagram with exact rows: δ H1(G, C) −−−−→ H0(G, A) −−−−→ H0(G, B) −−−−→ H0(G, C) −−−−→ 0       N 0  N 0  N 0   y Ay B y C y y 0 −−−−→ H0(G, A) −−−−→ H0(G, B) −−−−→ H0(G, C) −−−−→δ H1(G, A) In fact, the only places in doubt are the innermost and outermost squares. These are eas- ily checked directly, using the explicit definitions of cohomology and homology using a canonical resolution. Now we use the to get an exact sequence 0 0 0 0 0 0 KerNA → KerNB → KerNC → CokerNA → CokerNB → CokerNC , which is itself a functor of the exact sequence 0 → A → B → C → 0. It follows easily that we can add a H1(G, C) to the beginning of the sequence, and a H1(G, A) at the end. So we get a long two-sided exact sequence with the homology groups on the left, the snake lemma sequence in the middle, and the cohomology sequence on the right. All these are just the Tate groups though, just from the definitions. Hence: Theorem 1. Suppose 0 → A → B → C → 0 is a short exact sequence. Then there exists a long exact sequence · · · → Hˆ i(G, A) → Hˆ i(G, B) → Hˆ i(G, C) → Hˆ i+1(G, A) → Hˆ i+1(G, B) → .... So the Tate groups form a δ-functor of covariant functors. We show that it is universal. The next proposition shows this:

Proposition 3. Let A be an induced module, i.e. A = Z[G] ⊗Z B for an abelian group B. Then Hˆ i(G, A) = 0 for all i. In other words, A is cohomologically trivial.

Proof. Since G is finite, we see that A = HomZ(Z[G],B). Thus A is both induced and co- induced; hence the regular cohomology and homology groups for i > 0 vanish. Hence the only Tate groups we need consider are Hˆ 0(G, A) and Hˆ −1(G, A). These can be checked directly easily.  Restriction and corestriction extend to all the Tate groups, as is easy to see. Straightfor- ward computations direct from the definitions show that they are natural transformations of δ-functors. (The only place where thi sneeds to be checked is at i = −1, 0 where the cohomology and homology are glued together. This is straightforward.) We have Cor ◦ Res = (G : H) as before, by dimension-shifting or uniqueness. Proposition 4. Let n := (G : 1). Then the groups Hˆ i(G, A) have exponent dividing n.

Proof. Immediate from the above discussion, taking H := 1.  Proposition 5. If A is finitely generated, then the Tate groups are finite. GROUP COHOMOLOGY 7

Proof. Looking at the cochains (or chains) we see that the Tate groups are finitely gener- ated. By the structure theorem for finitely generated abelian groups, our assertion follows at once from the previous proposition. 

5. CUP-PRODUCTS Theorem 2. Let A, B be G-modules. Then there exist homomorphisms:

Hˆ p(G, A) × Hˆ q(G, B) → Hˆ p+q(G, A ⊗ B) for all p, q, which are functorial in A and B. They satisfy the following properties: (1) The map Hˆ 0(G, A)×Hˆ 0(G, B) → Hˆ 0(G, A⊗B) comes from the canonical map AG × BG → (A ⊗ B)G. (2) The map commutes with boundary homomorphisms in the following sense. Sup- pose 0 → A0 → A → A00 → 0 is an exact sequence, which remains exact when tensored with the G-module B. Then if a00 ∈ Hˆ p(G, A00), b ∈ Hˆ q(G, B), we have: (δa00).b = δ(a00.b). (3) Conversely if 0 → B0 → B → B00 → 0 is exact, and the sequence remains exact when tensored with A, we then have:

a.(δb00) = (−1)pδ(a.b00).

Proof. Cf. [?]. 

6. THE HERBRAND QUOTIENT Let G be a of order n, generated by s ∈ G. Then we define an exact complex C:

s−1 N s−1 N s−1 ... −−−−→ Z[G] −−−−→ Z[G] −−−−→ Z[G] −−−−→ Z[G] −−−−→ ..., where N is the norm homomorphism. Define the functor F by F (A) := H(A ⊗ C), i.e. the homology of A ⊗ C. This is immediately seen to be a δ-functor of A since the modules Z[G] are free, hence flat. The δ-functor also coincides with the Tate groups at 0, hence it consists of the Tate groups since both δ-functors are universal. In particular, the Tate groups are periodic with period (dividing) two. ˆ 1 Define the Herbrand quotient as h(A) := card(H (G,A)) . It is defined if both the nu- card(Hˆ 0(G,A)) merator and the denominator are finite groups. We can view the Herbrand quotient as an Euler-Poincare characteristic. Indeed, if 0 → A → B → C → 0 is a short exact sequence, then the usual long exact sequence becomes an exact hexagon, which I don’t know how to typeset. But from the exact hexagon it follows at once that the Herbrand quotient is multiplicative. Note that h(A) = 1 if A is finite. Indeed, we have an exact sequence

G 0 → A → A → A → A/IGA → 0, where the second map is multiplication by s − 1 for s a generator of G. We see that AG and AG have the same cardinality if A is finite. 8 AKHIL MATHEW

7. Let L/K be a finite Galois extension with group G. Then L is a G-module in the obvious manner. Note that K = LG by . Theorem 3 (Normal Basis Theorem). The additive group L is G-free. In other words, there exists a basis of L/K given by xσ, σ ∈ G, such that τxσ = xτσ for τ ∈ G. Proof. See [?].  By the previous section, we get: Theorem 4. L is cohomologically trivial over G. We next consider the multiplicative group L∗. It is also a G-module. We denote s(a) by as for s ∈ G, for convenience. Theorem 5 (Hilbert 90). H1(G, L∗) = 0. Proof. Let f : G → L∗ be a cocyle, i.e. f(ts) = f(t)f˙(s)t for all s, t ∈ G. Consider the sum X f(s)s ∈ L[G], s which is obviously a K-endomorphism of L. It cannot be identically zero, since distinct K- of L are linearly independent by a theorem of Dedekind. (See [?].) So pick a ∈ L∗ with: X b := f(s)as 6= 0. s Now consider bt: X X bt = f(s)tats = f(t)−1f(ts)ats = f(t)−1b, s s or in other words: b f(t) = . bt Hence f is a coboundary as well. 

8. COHOMOLOGICAL TRIVIALTY Let G be a finite group. We say that a G-module A is cohomologically trivial if all its Tate groups vanish. An induced module is cohomologically trivial. We now consider p-groups. Theorem 6. Let G be a p-group, A a G-module annihilated by p. If we have Hˆ i(G, A) = 0 for some i, then A is cohomologically trivial. Indeed A is free over Fp[G]. We need a lemma:

Lemma 1 (Noncommutative Nakayama). Suppose H0(G, A) = 0. Then A = 0. −1 0 Proof. We have Hˆ (G, A) = H0(G, A) = 0, i.e. A/IGA = 0. Consider now A := 0G Hom(A, Fp). We then have A = 0 since A = IGA and a G-homomorphism into Fp 0 must annihilate IGA, assuming of course that G acts trivially on Fp. Now A is a vector 0 space over Fp too. We show that A = 0, hence A = 0. Indeed, consider a finitely generated G-submodule of A0, say B0. If B0 = 0, this will establish our contention. Suppose B0 6= 0. But B0 is of order a power of p, and it is acted on by a group of order a power of p. Hence the number of fixed points is a power of p and thus consists not only of 0G the origin, so B 6= 0, contradiction.  GROUP COHOMOLOGY 9

As a simple corollary, if A = IGA, then A = 0. Alternatively, if A = IGA + B, then A = B. This is strikingly reminiscent of Nakayama’s lemma. of the theorem. Start with i = −2. We are given that H1(G, A) = 0. We will show that A is free over Fp[G]. Start by lifting a Fp-basis of A/IGA to A, say L = {u1, . . . , um}. Then L generates A by the above remarks. Let B be the free Fp[G]-module on L. We have a surjective G-homomorphism B → A. Let B0 be the kernel. Then we have an exact sequence 0 → B0 → B → A → 0, 0 Now by “dimension-shifting”, we get H0(G, B ) = H1(G, A) = 0. Indeed, we have an exact sequence 0 0 = H1(G, A) → H0(G, B ) → H0(G, B) → H0(G, A). 0 But B/IGB = A/IGA, so the last two are isomorphic. Thus H0(G, B ) = 0. Hence 0 B = 0, and B → A is an isomorphism. Hence A is free over Fp[G], proving the theorem in this case. Now suppose i 6= 2. Use a repeated dimension-shifting argument to get a module B whose Tate groups Hˆ i(H,B) for any H ⊂ G are a translation of those of A (i.e. Hˆ i(H,A)), such that the zero module for A falls at −2 for B. Then it follows by the above argument that B is Fp[G]-free, hence cohomologically trivial. Since A’s Tate groups (for all subgroups H) are just a translation, A is also cohomologically trivial.  We now consider the opposite scenario. Theorem 7. Let A be a G-module without p-torsion, G a p-group. Then A is cohomolog- ically trivial if and only if the Tate groups Hˆ i(G, A) vanish for consecutive indices. Proof. We have an exact sequence p 0 −−−−→ A −−−−→ A −−−−→ A/pA. Now A/pA has only p-torsion. If for two consecutive indices the Tate groups of A, i.e. Hˆ i(G, A) vanish, then a Tate group Hˆ i(G, A) of A/pA also vanishes by the exact se- quence of Tate groups. Hence A/pA is cohomologically trivial, so all its Tate groups Hˆ i(H, A/pA) = 0 for h ⊂ G and i ∈ Z. This implies again from the exact Tate group sequence that the maps p Hˆ i(H,A) −−−−→ Hˆ i(H,A) are isomorphisms. But Hˆ i(H,A) has only p-torsion. (Composing Cor and Res from the unit group gives a power of p.). Hence all the Tate groups of A vanish, proving the theorem. 

Theorem 8. Let A be a G-module. For each prime p, let Gp be a p-Sylow subgroup. np np+1 Suppose that for each p, we have Hˆ (Gp,A) = Hˆ (Gp,A) = 0. Then the Tate groups Hˆ i(G, A) = 0.

i Proof. We see immediately from the previous results that Hˆ (Gp,A) = 0 for all i. Now the composition Cor ◦ Res with respect to a subgroup Gp is just multiplication by a power of p, and it is zero. Hence the ideal I = {a : a annihilates the Tate groups Hˆ i(G, A) for all H, i contains a power of each prime, hence is the unit ideal, whence the theorem. In reality A is cohomologically trivial, but we won’t use this.  10 AKHIL MATHEW

9. TATE’S THEOREM Theorem 9. Let f : A → B be a G-homomorphism. Supposed that the induced homomor- phisms (still denoted by f) on the Tate groups are as follows: for each p, there is a p-Sylow np np subgroup Gp and an integer np such that f is surjective Hˆ (G, A) → Hˆ (G, A), bijec- tive for np + 1, and injective for np + 2. Then f induces isomorphisms on the cohomology groups. 0 ¯ Proof. Let A := Z[G] ⊗Z A be the standard induced module containing A. Let A : +A0 ⊕ B (direct sum). We have an injective map A → A¯ given by the canonical map direct summed with f. Let C be the cokernel. There is a short exact sequence 0 → A → A¯ → C → 0, which induces a long exact sequence · · · → Hˆ i(H,A) → Hˆ i(H, A¯) → Hˆ i(H,C) → Hˆ i+1(H,A) → ..., where H ⊂ G is a subgroup. Now the cohomology of A¯ is the same as that of B. Take H = Gp. Writing out the exact sequence and using the assumptions shows by earlier results that C is cohomologically trivial, whence the result.  Let G be a finite group. Let A be a G-module. We have a bilinear map A × Z → A, which is trivially a G-homomorphism (with G acting trivially on Z). Now let w ∈ Hˆ A(G, q). Cupping with w gives a morphism ˆ i ˆ i+q H (G, Z) → H (G, A ⊗Z Z), all i. In other words, we get a map: i i+q Hˆ (G, Z) → Hˆ (G, A). If q = 0, then this map just comes from Z → A ⊗ Z, which follows for i = 0 by the definitions of cup-products, and then by dimension-shifting. Theorem 10. Fix w as above. Suppose that for each p, the cup-product of Resw is in- np jective on Hˆ (Gp,A), bijective on the next group, and surjective on the one after that. (Same notation as before.) Then the cup-product induces isomorphisms from Hˆ i(G, Z) → Hˆ i+q(G, Z) for all i. For q = 0 this is the previous result. This can be seen by a dimension-shifting argument, and is of course plausible. See [?]. Now find an induced module A0 and an exact sequence 0 → A → A0 → A00 → 0. We then have boundary isomorphic maps Hˆ i(G, A00) → Hˆ i+1(G, A). Fix w now as before. Then δ(w.a00) = δw.a00. This is better explained on paper.