This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Algebraic structure on Tate-Hochschild cohomology of a Frobenius algebra

Satoshi Usui Graduate School of Mathematics, Tokyo University of Science, 1-3 Kagurazaka, Shinjuku-ku, Tokyo 162-8601 JAPAN [email protected]
Abstract.

We study cup product and cap product in Tate-Hochschild theory for a finite dimensional Frobenius algebra. We show that Tate-Hochschild cohomology ring equipped with cup product is isomorphic to singular Hochschild cohomology ring introduced by Wang.

An application of cap product occurs in Tate-Hochschild duality; as in Tate (co)homology of a finite group, the cap product with the fundamental class of a finite dimensional Frobenius algebra provides certain duality result between Tate-Hochschild cohomology and homology groups.

Moreover, we characterize minimal complete resolutions over a finite dimensional self-injective algebra by means of the notion of minimal complexes introduced by Avramov and Martsinkovsky.

Key words and phrases:
Tate-Hochschild (co)homology, Frobenius algebra, Cup product, Cap product
2010 Mathematics Subject Classification:
16E05, 16E40

Introduction

For an associative algebra AA which is projective over a ground commutative ring kk, Hochschild cohomology groups H(A,M)\operatorname{H}\nolimits^{*}(A,M) of AA with coefficients in an AA-bimodule MM are defined by the cohomology groups of the cochain complex 𝓂AkA(𝐏,M)\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A\otimes_{k}A^{\circ}}(\operatorname{\mathbf{P}}\nolimits,M), where 𝐏\operatorname{\mathbf{P}}\nolimits is arbitrary projective resolution of AA as an AA-bimoudle. Recall that Hi(A,M)\operatorname{H}\nolimits^{i}(A,M) is isomorphic to the space of morphisms from AA to 𝚺iM\boldsymbol{\Sigma}^{i}M in the bounded derived category 𝒟b(AkA)\mathcal{D}^{{\rm b}}(A\otimes_{k}A^{\circ}) of AA-bimodules for i0i\geq 0, where 𝚺\boldsymbol{\Sigma} is the shift functor. There is an operator, the so-called cup product \smile, H(A,M)H(A,N)H(A,MAN)\operatorname{H}\nolimits^{*}(A,M)\otimes\operatorname{H}\nolimits^{*}(A,N)\rightarrow\operatorname{H}\nolimits^{*}(A,M\otimes_{A}N) for AA-bimodules MM and NN. It is defined, on the level of complexes, by means of a diagonal approximation associated with a given projective resolution of AA. In particular, the cup product does not depend on the choice of a diagonal approximation and a projective resolution. If M=AM=A and N=AN=A, then Hochschild cohomology

H(A,A):=i0Hi(A,A)\operatorname{H}\nolimits^{\bullet}(A,A):=\bigoplus_{i\geq 0}\operatorname{H}\nolimits^{i}(A,A)

equipped with the cup product forms a graded algebra. In [Ger63], Gerstenhaber showed that the cup product induced by a certain diagonal approximation obtained from the bar resolution of an associative algebra is graded commutative, that is, the Hochschild cohomology ring based on the bar resolution is graded commutative. In conclusion, the result of him implies that the Hochschild cohomology ring via any projective resolution of an associative algebra is a graded commutative algebra.

Inspired by the Buchweitz’s result on Tate cohomology of Iwanaga-Gorenstein algebras ([Buc86]), Wang [Wan18] introduced singular Hochschild cochain complex Csg(A,A)C_{\rm sg}(A,A) of an associative algebra AA over a field and proved that the singular Hochschild cohomology group HHsgi(A,A)\operatorname{HH}\nolimits_{\rm sg}^{i}(A,A) of AA is isomorphic to the space of morphisms from AA to 𝚺iA\boldsymbol{\Sigma}^{i}A in the singularity category 𝒟sg(AkA)\mathcal{D}_{{\rm sg}}(A\otimes_{k}A^{\circ}) of AkAA\otimes_{k}A^{\circ} for any ii\in\operatorname{\mathbb{Z}}\nolimits. Furthermore, he discovered a differential graded associative and unital product on Csg(A,A)C_{\rm sg}(A,A) such that singular Hochschild cohomology

HHsg(A,A):=iHHsgi(A,A)\operatorname{HH}\nolimits_{\rm sg}^{\bullet}(A,A):=\bigoplus_{i\in\operatorname{\mathbb{Z}}\nolimits}\operatorname{HH}\nolimits_{\rm sg}^{i}(A,A)

equipped with the induced product is a graded commutative algebra.

In the case that an algebra AA is a finite dimensional Frobenius algebra, the singular Hochschild cohomology groups of AA coincide with the cohomology groups based on a complete resolution of AA. They are called Tate-Hochschild cohomology groups of AA and denoted by H^(A,A)\operatorname{\widehat{H}}\nolimits^{*}(A,A). Therefore, Tate-Hochschild cohomology

H^(A,A):=iH^i(A,A)\operatorname{\widehat{H}}\nolimits^{\bullet}(A,A):=\bigoplus_{i\in\operatorname{\mathbb{Z}}\nolimits}\operatorname{\widehat{H}}\nolimits^{i}(A,A)

becomes a graded commutative algebra whose structure depends on the singular Hochschild cohomology ring structure. On the other hand, Sanada [San92] constructed one (complete) diagonal approximation associated with the complete bar resolution of a finite dimensional Frobenius algebra. In particular, the cup product induced by his diagonal approximation makes the Tate-Hochschild cohomology ring into a graded commutative algebra. These results motivate us to ask the following questions:

  • (1)

    Is the Tate-Hochschild cohomology ring given by Sanada isomorphic to the singular Hochschild cohomology ring introduced by Wang?

  • (2)

    Is there the theory of cup product in Tate-Hochschild theory of any finite dimensional Frobenius algebra as in Hochschild theory?

Let us remark on the second question: Nguyen [Ngu13] has already developed the theory of cup product on Tate-Hochschild cohomology of a finite dimensional Hopf algebra.

In this paper, along the same lines as Brown [Bro94, Chapter VI, Section 5], we will develop the theory of cup product in Tate-Hochschild theory of a finite dimensional Frobenius algebra. More precisely, we will show that the existence of a diagonal approximation for arbitrary complete resolution of a given finite dimensional Frobenius algebra and that all diagonal approximations define exactly one cup product up to isomorphism (see Section 3.1). That is the answer to the question (2). Moreover, we will prove the following our main result, which is the answer to the question (1):

Main Theorem (Theorem 3.22).

Let AA be a finite dimensional Frobenius algebra over a field. Then there exists an isomorphism

H^(A,A)HHsg(A,A)\operatorname{\widehat{H}}\nolimits^{\bullet}(A,A)\cong\operatorname{HH}\nolimits_{{\rm sg}}^{\bullet}(A,A)

of graded commutative algebras.

We also deal with cap product in Tate-Hochschild theory and show that the cap product with the fundamental class of a finite dimensional Frobenius algebra gives certain duality result between Tate-Hochschild cohomology and homology groups. These results allow us to prove that the cup product on Tate-Hochschild cohomology contains not only the cup product, but also the cap product on Hochschild (co)homology.

Moreover, we provide a characterization of minimal complete resolutions of finitely generated modules over a finite dimensional self-injective algebra in the sense of Avramov and Martsinkovsky [AM02]. More concretely, we will show that any minimal complete resolution of a finitely generated module consists of its minimal projective resolution and its ((1-1)-shifted) minimal injective resolution.

This paper is organized as follows: in Section 1, we recall the basic notions related to Hochschild (co)homology groups and the cup product and the cap product on them. Section 2 is devoted to recalling the definitions of Tate and Tate-Hochschild (co)homology groups and to characterizing minimal complete resolutions over a finite dimensional self-injective algebra in terms of projective resolutions and injective resolutions. Section 3 contains our main theorem. Before proving it, we will define not only cup product, but also cap product by using a diagonal approximation, and we prove that these operators coincide with composition products. In Section 4, we will show that the cap product induces duality between Tate-Hochschild cohomology and homology groups. Using this result, we will prove that the cup product on Tate-Hochschild cohomology extends the cup product and the cap product on Hochschild (co)homology.

Throughout this paper, an algebra means an associative and unital algebra over a commutative ring, and all modules are left modules unless otherwise stated. The ground ring is taken to be a field when we assume a given algebra to be finite dimensional. For a kk-algebra AA, we denote D()=Homk(,k)D(-)=\operatorname{Hom}\nolimits_{k}(-,k) and ():=HomA(,A)(-)^{\vee}:=\operatorname{Hom}\nolimits_{A}(-,A). Finally, \otimes is an abbreviation for k\otimes_{k}.

1. Preliminaries

Following Brown [Bro94], we briefly recall some basic notions related to Hochschild (co)homology groups and the cup product and the cap product on them.

1.1. Hom complexes and tensor products of complexes

A chain complex C=(C,dC)C=(C,d^{C}) over an algebra AA is the pair of a graded AA-module C=nCnC=\bigoplus_{n\in\mathbb{Z}}C_{n} and a graded AA-linear map dC:CCd^{C}:C\rightarrow C of degree 1-1 such that (dC)2=0(d^{C})^{2}=0. Dually, a cochain complex C=(C,dC)C=(C,d_{C}) is the pair of a graded AA-module C=nCnC=\bigoplus_{n\in\mathbb{Z}}C^{n} and a graded AA-linear map dC:CCd_{C}:C\rightarrow C of degree 11 such that dC2=0d_{C}^{2}=0. In both cases, the graded map dd is called the differential of CC. Note that any cochain complex CC can be regarded as a chain complex by reindexing Cn=CnC_{n}=C^{-n}, and vice versa. For a chain complex CC and ii\in\mathbb{Z}, we define the ii-shifted chain complex 𝚺iC=(𝚺iC,d𝚺iC)\boldsymbol{\Sigma}^{i}C=(\boldsymbol{\Sigma}^{i}C,d^{\boldsymbol{\Sigma}^{i}C})

to be the chain complex given by (𝚺iC)n:=Cni(\boldsymbol{\Sigma}^{i}C)_{n}:=C_{n-i} and d𝚺iC:=(1)idCd^{\boldsymbol{\Sigma}^{i}C}:=(-1)^{i}d^{C}. For two chain complexes CC and CC^{\prime} of AA-modules, a chain map f:CCf:C\rightarrow C^{\prime} is a graded AA-linear map f:CCf:C\rightarrow C^{\prime} of degree 0 such that dCf=fdCd^{C^{\prime}}f=fd^{C}.

For two chain maps f,g:CCf,g:C\rightarrow C^{\prime}, we say that ff is homotopic to gg if there exists a graded AA-linear map h:CCh:C\rightarrow C^{\prime} of degree 11 such that fg=dCh+hdCf-g=d^{C^{\prime}}h+hd^{C}. Then the graded map hh is called a homotopy from ff to gg. In such a case, we denote fgf\sim g. A chain map f:CCf:C\rightarrow C^{\prime} is called null-homotopic if f0f\sim 0. It is well-known that the relation \sim is an equivalence relation. We denote by [C,C][C,C^{\prime}] the quotient kk-module induced by this equivalence relation. A chain map f:CCf:C\rightarrow C^{\prime} is a homotopy equivalence if there exists a chain map g:CCg:C^{\prime}\rightarrow C such that fgidCfg\sim\operatorname{id}\nolimits_{C^{\prime}} and gfidCgf\sim\operatorname{id}\nolimits_{C}, and a chain map f:CCf:C\rightarrow C^{\prime} is a quasi-isomorphism if the morphism Hn(f):Hn(C)Hn(C)\operatorname{H}\nolimits_{n}(f):\operatorname{H}\nolimits_{n}(C)\rightarrow\operatorname{H}\nolimits_{n}(C^{\prime}) induced by ff is an isomorphism for any nn\in\operatorname{\mathbb{Z}}\nolimits, where Hn(C)\operatorname{H}\nolimits_{n}(C) is the nn-th homology group of the complex CC defined by the quotient AA-module Zn(C)/Bn(C)Z_{n}(C)/B_{n}(C) with Zn(C)=KerdnCZ_{n}(C)=\operatorname{Ker}\nolimits d^{C}_{n} and Bn(C)=Imdn+1CB_{n}(C)=\operatorname{Im}\nolimits d^{C}_{n+1}. We say that a chain complex CC is acyclic if Hn(C)=0\operatorname{H}\nolimits_{n}(C)=0 for every nn\in\operatorname{\mathbb{Z}}\nolimits and that a chain complex CC is contractible if idC0id_{C}\sim 0. Then a homotopy from idC\operatorname{id}\nolimits_{C} to 0 is called a contracting homotopy.

The following easy and well-known lemma and its dual lemma play important roles in this paper. We will include a proof of the first lemma.

Lemma 1.1 ([Bro94, Chapter I, Lemma 7.4]).

Let CC and CC^{\prime} be chain complexes of AA-modules and rr\in\operatorname{\mathbb{Z}}\nolimits. Suppose that CiC_{i} is projective over AA for i>ri>r and that Hi(C)=0\operatorname{H}\nolimits_{i}(C^{\prime})=0 for iri\geq r. Then

  1. (1)

    Any family {fi:CiCi}ir\{f_{i}:C_{i}\rightarrow C^{\prime}_{i}\}_{i\leq r} commuting with differentials extends to a chain map f:CCf:C\rightarrow C^{\prime}.

  2. (2)

    Let f,g:CCf,g:C\rightarrow C^{\prime} be chain maps and {hi:CiCi+1}ir\{h_{i}:C_{i}\rightarrow C^{\prime}_{i+1}\}_{i\leq r} a family of AA-linear maps such that di+1Chi+hi1diC=figid^{C^{\prime}}_{i+1}h_{i}+h_{i-1}d^{C}_{i}=f_{i}-g_{i} for iri\geq r. Then {hi}ir\{h_{i}\}_{i\leq r} extends to a homotopy from ff to gg.

Proof.

We only show (1)(1); the proof of (2)(2) is similar. For each rr\in\operatorname{\mathbb{Z}}\nolimits, let drC=ιrπrd_{r}^{C^{\prime}}=\iota_{r}^{\prime}\pi_{r}^{\prime} be the canonical factorization through ImdrC\operatorname{Im}\nolimits d_{r}^{C^{\prime}}. Since we have drCfrdr+1C=0d_{r}^{C^{\prime}}f_{r}d_{r+1}^{C}=0 and Hr(C)=0\operatorname{H}\nolimits_{r}(C^{\prime})=0, there exists a lifting morphism fr+1f_{r+1}^{\prime} of frdr+1Cf_{r}d_{r+1}^{C} such that frdr+1C=ιrfr+1f_{r}d_{r+1}^{C}=\iota^{\prime}_{r}f_{r+1}^{\prime}. The projectivity of Cr+1C_{r+1} implies that there exists a morphism fr+1:Cr+1Cr+1f_{r+1}:C_{r+1}\rightarrow C_{r+1}^{\prime} such that fr+1=πr+1fr+1f^{\prime}_{r+1}=\pi_{r+1}^{\prime}f_{r+1}, so that we have

dr+1Cfr+1=ιr(πr+1fr+1)=ιrfr+1=frdr+1C.\displaystyle d_{r+1}^{C^{\prime}}f_{r+1}=\iota^{\prime}_{r}(\pi_{r+1}^{\prime}f_{r+1})=\iota^{\prime}_{r}f^{\prime}_{r+1}=f_{r}d_{r+1}^{C}.

Repeating this argument inductively, we obtain the desired chain map. ∎

Lemma 1.2.

Let CC and CC^{\prime} be chain complexes of AA-modules and rr\in\operatorname{\mathbb{Z}}\nolimits. Suppose that CiC^{\prime}_{i} is injective over AA for i<ri<r and that Hi(C)=0\operatorname{H}\nolimits_{i}(C)=0 for iri\leq r. Then

  1. (1)

    Any family {fi:CiCi}ir\{f_{i}:C_{i}\rightarrow C^{\prime}_{i}\}_{i\geq r} of morphisms of AA-modules commuting with differentials extends to a chain map f:CCf:C\rightarrow C^{\prime}.

  2. (2)

    Let f,g:CCf,g:C\rightarrow C^{\prime} be chain maps and {hi:CiCi+1}ir\{h_{i}:C_{i}\rightarrow C^{\prime}_{i+1}\}_{i\geq r} a family of AA-linear maps such that di+1Chi+hi1diC=figid^{C^{\prime}}_{i+1}h_{i}+h_{i-1}d^{C}_{i}=f_{i}-g_{i} for iri\geq r. Then {hi}ir\{h_{i}\}_{i\geq r} extends to a homotopy from ff to gg.

Let CC and CC^{\prime} be two chain complexes of AA-modules. We define the tensor product CACC\otimes_{A}C^{\prime} as the chain complex with components

(CAC)n:=iCniACi\displaystyle(C\otimes_{A}C^{\prime})_{n}:=\bigoplus_{i\in\operatorname{\mathbb{Z}}\nolimits}C_{n-i}\otimes_{A}C^{\prime}_{i}

and differentials dnCAC:(CAC)n(CAC)n1d^{C\otimes_{A}C^{\prime}}_{n}:(C\otimes_{A}C^{\prime})_{n}\rightarrow(C\otimes_{A}C^{\prime})_{n-1} given by

dnCAC(xAx):=dC(x)x+(1)|x|xdC(x)\displaystyle d^{C\otimes_{A}C^{\prime}}_{n}(x\otimes_{A}x^{\prime}):=d^{C}(x)\otimes x^{\prime}+(-1)^{|x|}x\otimes d^{C^{\prime}}(x^{\prime})

for homogeneous elements xCx\in C and xCx^{\prime}\in C^{\prime}.

Now we define 𝓂A(C,C)\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A}(C,C^{\prime}) to be the chain complex with components

𝓂A(C,C)n:=iZHomA(Ci,Ci+n)\displaystyle\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A}(C,C^{\prime})_{n}:=\prod_{i\in Z}\operatorname{Hom}\nolimits_{A}(C_{i},C^{\prime}_{i+n})

and differentials

dn𝓂A(C,C):𝓂A(C,C)n𝓂A(C,C)n1d^{\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A}(C,C^{\prime})}_{n}:\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A}(C,C^{\prime})_{n}\rightarrow\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A}(C,C^{\prime})_{n-1}

given by

dn𝓂A(C,C)(f):=(di+nCfi(1)nfi1diC)i\displaystyle d^{\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A}(C,C^{\prime})}_{n}(f):=\left(d^{C^{\prime}}_{i+n}f_{i}-(-1)^{n}f_{i-1}d^{C}_{i}\right)_{i\in\operatorname{\mathbb{Z}}\nolimits}

for any f=(fi)i𝓂A(C,C)nf=(f_{i})_{i\in\operatorname{\mathbb{Z}}\nolimits}\in\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A}(C,C^{\prime})_{n}. One sees that Hn(𝓂A(C,C))=[C,𝚺nC]\operatorname{H}\nolimits_{n}(\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A}(C,C^{\prime}))=[C,\boldsymbol{\Sigma}^{-n}C^{\prime}] for any nn\in\operatorname{\mathbb{Z}}\nolimits. We put [C,C]n:=[C,𝚺nC][C,C^{\prime}]_{n}:=[C,\boldsymbol{\Sigma}^{-n}C^{\prime}].

Let CC be a chain complex of left AA-modules. We define the kk-dual complex 𝔻(C)\operatorname{\mathbb{D}}\nolimits(C) of CC by the cochain complex 𝓂k(C,k)\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{k}(C,k) of right AA-modules. Moreover, the AA-dual complex CC^{\vee} of CC is defined as the cochain complex 𝓂A(C,A)\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A}(C,A) of right AA-modules.

Let C,CC,C^{\prime} and C′′C^{\prime\prime} be chain complexes of AA-modules. The composition of graded maps is defined to be the chain map

𝓂A(C,C′′)𝓂A(C,C)𝓂A(C,C′′)\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A}(C^{\prime},C^{\prime\prime})\otimes\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A}(C,C^{\prime})\xrightarrow{\circ}\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A}(C,C^{\prime\prime})

sending uv𝓂A(C,C′′)s𝓂A(C,C)ru\otimes v\in\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A}(C^{\prime},C^{\prime\prime})_{s}\otimes\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A}(C,C^{\prime})_{r} to

uv:=(up+rvp)p𝓂A(C,C′′)r+s.u\circ v:=(u_{p+r}v_{p})_{p\in\operatorname{\mathbb{Z}}\nolimits}\in\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A}(C,C^{\prime\prime})_{r+s}.

Hence the chain map induces a well-defined operator

H(𝓂A(C,C′′))H(𝓂A(C,C))H(𝓂A(C,C′′)).\operatorname{H}\nolimits_{*}(\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A}(C^{\prime},C^{\prime\prime}))\otimes\operatorname{H}\nolimits_{*}(\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A}(C,C^{\prime}))\rightarrow\operatorname{H}\nolimits_{*}(\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A}(C,C^{\prime\prime})).

Let C1,C2,C1C_{1},C_{2},C^{\prime}_{1} and C2C^{\prime}_{2} be chain complexes of AA-modules, and let u:C1C1u:C_{1}\rightarrow C^{\prime}_{1} and v:C2C2v:C_{2}\rightarrow C^{\prime}_{2} be graded maps of degree ss and of degree tt, respectively. The tensor product uAvu\otimes_{A}v of uu and vv is defined as the graded map of degree s+ts+t given by

(uAv)(x1x2):=(1)|v||x1|u(x1)Av(x2)(u\otimes_{A}v)(x_{1}\otimes x_{2}):=(-1)^{|v||x_{1}|}u(x_{1})\otimes_{A}v(x_{2})

for homogeneous elements x1C1x_{1}\in C_{1} and x2C2x_{2}\in C_{2}. The tensor product of graded maps gives rise to a chain map

𝓂A(C1,C1)𝓂A(C2,C2)𝓂A(C1AC2,C1AC2).\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A}(C_{1},C^{\prime}_{1})\otimes\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A}(C_{2},C^{\prime}_{2})\rightarrow\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A}(C_{1}\otimes_{A}C_{2},C^{\prime}_{1}\otimes_{A}C^{\prime}_{2}).

1.2. Hochschild (co)homology groups

Throughout this section and the next, let AA be a kk-algebra which is projective over kk. Let AeA^{\textrm{e}} denote the enveloping algebra AAA\otimes A^{\circ} of AA, where AA^{\circ} is the opposite algebra of AA. Suppose that all projective resolutions of AA are taken over AeA^{\textrm{e}}. We view any module MM as a complex concentrated in degree 0. Let 𝐏=i0Pi\operatorname{\mathbf{P}}\nolimits=\bigoplus_{i\geq 0}P_{i} be a projective resolution of AA, that is, a quasi-isomorphism 𝐏𝜀A\operatorname{\mathbf{P}}\nolimits\xrightarrow{\varepsilon}A with PiP_{i} finitely generated projective. The epimorphism ε:P0A\varepsilon:P_{0}\rightarrow A is called the augmentation of 𝐏\operatorname{\mathbf{P}}\nolimits.

The nn-th Hochschild cohomology group Hn(A,M)\operatorname{H}\nolimits^{n}(A,M) of AA with coefficients in an AA-bimodule MM is defined by Hn(A,M):=Hn(𝓂Ae(𝐏,M))\operatorname{H}\nolimits^{n}(A,M):=\operatorname{H}\nolimits^{n}(\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{P}}\nolimits,M)), where the nn-th component of 𝓂Ae(𝐏,M)\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{P}}\nolimits,M) is given by

𝓂Ae(𝐏,M)n=𝓂Ae(𝐏,M)n=HomAe(Pn,M).\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{P}}\nolimits,M)^{n}=\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{P}}\nolimits,M)_{-n}=\operatorname{Hom}\nolimits_{A^{\textrm{e}}}(P_{n},M).

The nn-th Hochschild homology group Hn(A,M)\operatorname{H}\nolimits_{n}(A,M) of AA with coefficients in an AA-bimodule MM is defined as Hn(A,M):=Hn(𝐏AeM)\operatorname{H}\nolimits_{n}(A,M):=\operatorname{H}\nolimits_{n}(\operatorname{\mathbf{P}}\nolimits\otimes_{A^{\textrm{e}}}M).

1.3. Cup product and cap product in Hochschild theory

In this section, we recall the definitions of the cup product and of the cap product in Hochschild theory. If 𝐏𝜀A\operatorname{\mathbf{P}}\nolimits\xrightarrow{\varepsilon}A is a projective resolution, we can associate 𝐏\operatorname{\mathbf{P}}\nolimits with the following augmented chain complex having AA in degree 1-1:

P2P1P0𝜀A00.\cdots\rightarrow P_{2}\rightarrow P_{1}\rightarrow P_{0}\xrightarrow{\varepsilon}A\rightarrow 0\rightarrow 0\rightarrow\cdots.

For two projective resolutions 𝐏𝜀A\operatorname{\mathbf{P}}\nolimits\xrightarrow{\varepsilon}A and 𝐏εA\operatorname{\mathbf{P}}\nolimits^{\prime}\xrightarrow{\varepsilon^{\prime}}A, a chain map f:𝐏𝐏f:\operatorname{\mathbf{P}}\nolimits\rightarrow\operatorname{\mathbf{P}}\nolimits^{\prime} is called an augmentation-preserving chain map if εf0=ε\varepsilon^{\prime}f_{0}=\varepsilon. Since the tensor product 𝐏A𝐏\operatorname{\mathbf{P}}\nolimits\otimes_{A}\operatorname{\mathbf{P}}\nolimits of 𝐏\operatorname{\mathbf{P}}\nolimits with itself is also a projective resolution of AA with augmentation εAε:P0AP0A\varepsilon\otimes_{A}\varepsilon:P_{0}\otimes_{A}P_{0}\rightarrow A, there exists an augmentation-preserving chain map Δ:𝐏𝐏A𝐏\Delta:\operatorname{\mathbf{P}}\nolimits\rightarrow\operatorname{\mathbf{P}}\nolimits\otimes_{A}\operatorname{\mathbf{P}}\nolimits. We call such a chain map a diagonal approximation. For any AA-bimodules MM and NN, we define a graded kk-linear map

:𝓂Ae(𝐏,M)𝓂Ae(𝐏,N)𝓂Ae(𝐏,MAN)\smile:\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{P}}\nolimits,M)\otimes\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{P}}\nolimits,N)\rightarrow\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{P}}\nolimits,M\otimes_{A}N)

by

uv:=(uAv)Δu\smile v:=(u\otimes_{A}v)\Delta

for homogeneous elements u𝓂Ae(𝐏,M)u\in\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{P}}\nolimits,M) and v𝓂Ae(𝐏,N)v\in\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{P}}\nolimits,N). Since the diagonal approximation Δ\Delta is a chain map, we see that the map \smile is a chain map. Thus it induces a well-defined operator

:H(A,M)H(A,N)H(A,MAN).\smile:\operatorname{H}\nolimits^{*}(A,M)\otimes\operatorname{H}\nolimits^{*}(A,N)\rightarrow\operatorname{H}\nolimits^{*}(A,M\otimes_{A}N).

For uHm(A,M)u\in\operatorname{H}\nolimits^{m}(A,M) and vHn(A,N)v\in\operatorname{H}\nolimits^{n}(A,N), we call uvHm+n(A,MAN)u\smile v\in\operatorname{H}\nolimits^{m+n}(A,M\otimes_{A}N) the cup product of uu and vv.

Example 1.3.

Let AA be an algebra over a field kk, and let A¯\overline{A} denote a quotient vector space A/(k1)A/(k\cdot 1). For simplicity, we write a¯1,n\overline{a}_{1,n} for a¯1a¯nA¯n\overline{a}_{1}\otimes\cdots\otimes\overline{a}_{n}\in\overline{A}^{\otimes n}. The normalized bar resolution Bar(A)\mathrm{Bar}(A) of AA is the chain complex with Bar(A)n:=AA¯nA\mathrm{Bar}(A)_{n}:=A\otimes\overline{A}^{\otimes n}\otimes A and differentials

dnBar(A)(a0a¯1,nan+1)=i=0n(1)ia0a¯1,i1aiai+1¯a¯i+1,nan+1.d_{n}^{\mathrm{Bar}(A)}(a_{0}\otimes\overline{a}_{1,n}\otimes a_{n+1})=\sum_{i=0}^{n}(-1)^{i}a_{0}\otimes\overline{a}_{1,i-1}\otimes\overline{a_{i}a_{i+1}}\otimes\overline{a}_{i+1,n}\otimes a_{n+1}.

Then Bar(A)\mathrm{Bar}(A) is a projective resolution of AA. It is known that the graded AeA^{\textrm{e}}-linear map Δ:Bar(A)Bar(A)ABar(A)\Delta:\mathrm{Bar}(A)\rightarrow\mathrm{Bar}(A)\otimes_{A}\mathrm{Bar}(A) given by

Δ(a0\displaystyle\Delta(a_{0}\otimes a¯1,nan+1):=i=0n(a0a¯1,i1)A(1a¯i+1,nan+1)\displaystyle\,\overline{a}_{1,n}\otimes a_{n+1}):=\sum_{i=0}^{n}(a_{0}\otimes\overline{a}_{1,i}\otimes 1)\otimes_{A}(1\otimes\overline{a}_{i+1,n}\otimes a_{n+1})

is a chain map. Then we see that

(uv)\displaystyle(u\smile v) (a0a¯1,m+nam+n+1)\displaystyle(a_{0}\otimes\overline{a}_{1,m+n}\otimes a_{m+n+1})
=(1)mnu(a0a¯1,m1)Av(1a¯m,m+nam+n+1)\displaystyle=(-1)^{mn}u(a_{0}\otimes\overline{a}_{1,m}\otimes 1)\otimes_{A}v(1\otimes\overline{a}_{m,m+n}\otimes a_{m+n+1})

for u𝓂Ae(Bar(A),M)mu\in\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\mathrm{Bar}(A),M)^{m} and v𝓂Ae(Bar(A),N)nv\in\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\mathrm{Bar}(A),N)^{n}.

Consider the chain map

𝓂Ae(𝐏,M)𝓂Ae(𝐏,𝐏AN)𝓂Ae(𝐏,MAN)\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{P}}\nolimits,M)\otimes\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{P}}\nolimits,\operatorname{\mathbf{P}}\nolimits\otimes_{A}N)\rightarrow\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{P}}\nolimits,M\otimes_{A}N)

defined by uv(uAidM)vu\otimes v\mapsto(u\otimes_{A}\operatorname{id}\nolimits_{M})v for homogeneous elements u𝓂Ae(𝐏,M)u\in\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{P}}\nolimits,M) and v𝓂Ae(𝐏,𝐏AN)v\in\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{P}}\nolimits,\operatorname{\mathbf{P}}\nolimits\otimes_{A}N). Then it induces a well-defined operator, called the composition product,

(1.1) H(A,M)H(𝓂Ae(𝐏,𝐏AN))H(A,MAN).\displaystyle\operatorname{H}\nolimits^{*}(A,M)\otimes\operatorname{H}\nolimits^{*}(\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{P}}\nolimits,\operatorname{\mathbf{P}}\nolimits\otimes_{A}N))\rightarrow\operatorname{H}\nolimits^{*}(A,M\otimes_{A}N).

Since AA is projective as a right AA-module, the augmented chain complex 𝐏𝜀A\operatorname{\mathbf{P}}\nolimits\xrightarrow{\varepsilon}A is contractible as a complex of right AA-modules, so that the tensor product 𝐏ANAAN\operatorname{\mathbf{P}}\nolimits\otimes_{A}N\rightarrow A\otimes_{A}N is acyclic, which means that εAidN:𝐏ANN\varepsilon\otimes_{A}\operatorname{id}\nolimits_{N}:\operatorname{\mathbf{P}}\nolimits\otimes_{A}N\rightarrow N is a quasi-isomorphism. It follows from [Bro94, Chapter I, Theorem 8.5] that εAidN\varepsilon\otimes_{A}\operatorname{id}\nolimits_{N} induces an isomorphism

Hn(A,N)Hn(𝓂Ae(𝐏,𝐏AN)).\displaystyle\operatorname{H}\nolimits^{n}(A,N)\cong\operatorname{H}\nolimits^{n}(\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{P}}\nolimits,\operatorname{\mathbf{P}}\nolimits\otimes_{A}N)).
Theorem 1.4 ([BGSS08, Proposition 1.1]).

The cup product

:H(A,M)H(A,N)H(A,MAN).\smile:\operatorname{H}\nolimits^{*}(A,M)\otimes\operatorname{H}\nolimits^{*}(A,N)\rightarrow\operatorname{H}\nolimits^{*}(A,M\otimes_{A}N).

coincides with the composition product (1.1)(\ref{eq:15}) via the isomorphism above.

We now recall the definition of the cap product between Hochschild cohomology and homology groups and the statement analogue to Theorem 1.4. Consider the chain map

γ:𝓂Ae(𝐏,M)((𝐏A𝐏)AeN)𝐏Ae(MAN)\gamma:\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{P}}\nolimits,M)\otimes((\operatorname{\mathbf{P}}\nolimits\otimes_{A}\operatorname{\mathbf{P}}\nolimits)\otimes_{A^{\textrm{e}}}N)\rightarrow\operatorname{\mathbf{P}}\nolimits\otimes_{A^{\textrm{e}}}(M\otimes_{A}N)

given by

γ(uxAyAen):=(1)|u||x|xAeu(y)An,\displaystyle\gamma(u\otimes x\otimes_{A}y\otimes_{A^{\textrm{e}}}n):=(-1)^{|u||x|}x\otimes_{A^{\textrm{e}}}u(y)\otimes_{A}n,

where u𝓂Ae(𝐏,M)u\in\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{P}}\nolimits,M) and nAexAy((𝐏A𝐏)AeN)n\otimes_{A^{\textrm{e}}}x\otimes_{A}y\in((\operatorname{\mathbf{P}}\nolimits\otimes_{A}\operatorname{\mathbf{P}}\nolimits)\otimes_{A^{\textrm{e}}}N) are homogeneous elements and we use an isomorphism

(𝐏AM)AeN𝐏Ae(MAN).\displaystyle(\operatorname{\mathbf{P}}\nolimits\otimes_{A}M)\otimes_{A^{\textrm{e}}}N\cong\operatorname{\mathbf{P}}\nolimits\otimes_{A^{\textrm{e}}}(M\otimes_{A}N).

Then the chain map

:𝓂Ae(𝐏,M)(𝐏AeN)𝐏Ae(MAN)\frown:\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{P}}\nolimits,M)\otimes(\operatorname{\mathbf{P}}\nolimits\otimes_{A^{\textrm{e}}}N)\rightarrow\operatorname{\mathbf{P}}\nolimits\otimes_{A^{\textrm{e}}}(M\otimes_{A}N)

defined to be the composition of two chain maps γ\gamma and

id\displaystyle\operatorname{id}\nolimits (ΔAeid):\displaystyle\otimes(\Delta\otimes_{A^{\textrm{e}}}\operatorname{id}\nolimits):
𝓂Ae(𝐏,M)(𝐏AeN)𝓂Ae(𝐏,M)((𝐏A𝐏)AeN)\displaystyle\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{P}}\nolimits,M)\otimes(\operatorname{\mathbf{P}}\nolimits\otimes_{A^{\textrm{e}}}N)\rightarrow\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{P}}\nolimits,M)\otimes((\operatorname{\mathbf{P}}\nolimits\otimes_{A}\operatorname{\mathbf{P}}\nolimits)\otimes_{A^{\textrm{e}}}N)

gives rise to a well-defined operator

:Hm(A,M)Hn(A,N)Hnm(A,MAN).\frown:\operatorname{H}\nolimits^{m}(A,M)\otimes\operatorname{H}\nolimits_{n}(A,N)\rightarrow\operatorname{H}\nolimits_{n-m}(A,M\otimes_{A}N).

For uHm(A,M)u\in\operatorname{H}\nolimits^{m}(A,M) and wHn(A,N)w\in\operatorname{H}\nolimits_{n}(A,N), we call uwHnm(A,MAN)u\frown w\in\operatorname{H}\nolimits_{n-m}(A,M\otimes_{A}N) the cap product of uu and ww.

Example 1.5.

Using the same projective resolution and diagonal approximation as in Example 1.3, we see that

uw=(1)m(pm)(a0a¯1,pm1)Aeu(1a¯pm+1,p+1)An,\displaystyle u\frown w=(-1)^{m(p-m)}(a_{0}\otimes\overline{a}_{1,p-m}\otimes 1)\otimes_{A^{\textrm{e}}}u(1\otimes\overline{a}_{p-m+1,p+1})\otimes_{A}n,

for w=(a0a¯1,pap+1)Aen(Bar(A)AeN)pw=(a_{0}\otimes\overline{a}_{1,p}\otimes a_{p+1})\otimes_{A^{\textrm{e}}}n\in(\mathrm{Bar}(A)\otimes_{A^{\textrm{e}}}N)_{p} and u𝓂Ae(Bar(A),M)mu\in\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\mathrm{Bar}(A),M)^{m} with pm0p-m\geq 0.

On the other hand, we define a chain map

(1.2) 𝓂Ae(𝐏,M)(𝐏Ae(𝐏AN))𝐏Ae(MAN)\displaystyle\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{P}}\nolimits,M)\otimes(\operatorname{\mathbf{P}}\nolimits\otimes_{A^{\textrm{e}}}(\operatorname{\mathbf{P}}\nolimits\otimes_{A}N))\rightarrow\operatorname{\mathbf{P}}\nolimits\otimes_{A^{\textrm{e}}}(M\otimes_{A}N)

by

u(xAe(yAn))(1)|u||x|xAe(u(y)An)u\otimes(x\otimes_{A^{\textrm{e}}}(y\otimes_{A}n))\mapsto(-1)^{|u||x|}x\otimes_{A^{\textrm{e}}}(u(y)\otimes_{A}n)

for homogeneous elements u𝓂Ae(𝐏,M)u\in\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{P}}\nolimits,M) and (xAy)Aen(𝐏A𝐏)AeN(x\otimes_{A}y)\otimes_{A^{\textrm{e}}}n\in(\operatorname{\mathbf{P}}\nolimits\otimes_{A}\operatorname{\mathbf{P}}\nolimits)\otimes_{A^{\textrm{e}}}N. Moreover, it follows from [Bro94, Chapter I, Theorem 8.6] that there exists an isomorphism

(1.3) H(A,N)H(𝐏Ae(𝐏AN)).\displaystyle\operatorname{H}\nolimits_{*}(A,N)\cong\operatorname{H}\nolimits_{*}(\operatorname{\mathbf{P}}\nolimits\otimes_{A^{\textrm{e}}}(\operatorname{\mathbf{P}}\nolimits\otimes_{A}N)).

One proves the following in a similar way to the proof of Theorem 3.15(2)\ref{theo:3}(2) below.

Theorem 1.6.

The cap product

:H(A,M)H(A,N)H(A,MAN).\frown:\operatorname{H}\nolimits^{*}(A,M)\otimes\operatorname{H}\nolimits_{*}(A,N)\rightarrow\operatorname{H}\nolimits_{*}(A,M\otimes_{A}N).

agrees with the product induced by the chain map (1.2)(\ref{eq:1}) via the isomorphism (1.3)(\ref{eq:14}).

Remark that the case for M=N=AM=N=A is proved by Armenta [Arm19, Section 4].

2. Tate-Hochschild (co)homology groups of a self-injective algebra

Our aim in this section is to recall the definitions of complete resolutions over finite dimensional self-injective algebras and of Tate and Tate-Hochschild (co)homology groups. Moreover, we provide a characterization of minimal complete resolutions. Throughout this section, assume that AA is a finite dimensional algebra over a field kk.

2.1. Twisted bimodules

Let us begin with the preparation for some notation. We denote by Aut(A)\operatorname{Aut}\nolimits(A) the group of algebra automorphisms of AA. Note that any αAut(A)\alpha\in\operatorname{Aut}\nolimits(A) gives rise to αAut(A)\alpha^{\circ}\in\operatorname{Aut}\nolimits(A^{\circ}) defined by α(a):=α(a)\alpha^{\circ}(a^{\circ}):=\alpha(a)^{\circ} for any aAa^{\circ}\in A^{\circ}. For an AA-bimodule MM and two automorphisms α,βAut(A)\alpha,\beta\in\operatorname{Aut}\nolimits(A), we denote by Mβα{}_{\alpha}M_{\beta} the AA-bimodule which is MM as a kk-module and whose AA-bimodule structure is given by amb:=α(a)mβ(b)a*m*b:=\alpha(a)m\beta(b) for a,bAa,b\in A and mMβαm\in{}_{\alpha}M_{\beta}. We denote Mβ1:=Mβid{}_{1}M_{\beta}:={}_{\operatorname{id}\nolimits}M_{\beta} and M1α:=Midα{}_{\alpha}M_{1}:={}_{\alpha}M_{\operatorname{id}\nolimits}. Recall that we can identify an AA-bimodule MM with the left (right) AeA^{\textrm{e}}-module MM of which the structure is defined by (ab)m:=amb(m(ab):=bma)(a\otimes b^{\circ})m:=amb\ (m(a\otimes b^{\circ}):=bma) for abAea\otimes b^{\circ}\in A^{\textrm{e}} and mMm\in M. Using this identification, we have Mαβ=Mβα=Mβα{}_{\alpha\otimes\beta}M={}_{\alpha}M_{\beta}=M_{\beta\otimes\alpha}, where we set αβ:=αβ\alpha\otimes\beta:=\alpha\otimes\beta^{\circ} and βα=:βα\beta\otimes\alpha=:\beta\otimes\alpha^{\circ}. For any morphism f:MNf:M\rightarrow N of AA-bimodules and α\alpha, βAut(A)\beta\in\operatorname{Aut}\nolimits(A), there exists an isomorphism

HomAe(M,N)HomAe(Mβα,Nβα);ff,\displaystyle\operatorname{Hom}\nolimits_{A^{\textrm{e}}}(M,N)\rightarrow\operatorname{Hom}\nolimits_{A^{\textrm{e}}}({}_{\alpha}M_{\beta},{}_{\alpha}N_{\beta});\ f\mapsto f,

which is natural in both MM and NN.

2.2. Self-injective algebras and Frobenius algebras

In this subsection, we recall the definitions of self-injective algebras and of Frobenius algebras. Recall that a finite dimensional algebra AA is a self-injective algebra if AA is injective as a left or as a right AA-module, or equivalently, AA is injective as a left and as a right AA-module. Moreover, recall that a finite dimensional kk-algebra AA is a Frobenius algebra if there exists a non-degenerate bilinear form ,:AAk\langle-,-\rangle:A\otimes A\rightarrow k satisfying ab,c=a,bc\langle ab,c\rangle=\langle a,bc\rangle for a,ba,b and cAc\in A. The bilinear form gives rise to a left and a right AA-module isomorphism

(2.1) t1:AADA(A);x,x,\displaystyle t_{1}:{}_{A}A\rightarrow{}_{A}D(A);\ x\mapsto\langle-,x\rangle,
AAD(A)A;xx,,\displaystyle A_{A}\rightarrow D(A)_{A};\ x\mapsto\langle x,-\rangle,

where the left and the right AA-module structure of D(A)D(A) is given by

(af)(x):=f(xa) and (fa)(x):=f(ax)\displaystyle(af)(x):=f(xa)\mbox{ and }(fa)(x):=f(ax)

for a,xAa,x\in A and fD(A)f\in D(A). In particular, a Frobenius algebra is a self-injective algebra. Let {u1,,ur}\{u_{1},\ldots,u_{r}\} be a kk-basis of AA. Then we have another kk-basis {v1,,vr}\{v_{1},\ldots,v_{r}\} such that vi,uj=δij\langle v_{i},u_{j}\rangle=\delta_{ij} for all 1i,jr1\leq i,j\leq r, where δij\delta_{ij} denotes the Kronecker delta. We call {vi}i\{v_{i}\}_{i} the dual basis of {ui}i\{u_{i}\}_{i}. It is known that there exists an algebra automorphism ν\nu of AA such that a,b=b,ν(a)\langle a,b\rangle=\langle b,\nu(a)\rangle for a,bAa,b\in A. The automorphism ν\nu is unique, up to inner automorphism, and we call it the Nakayama automorphism of AA. The Nakayama automorphism ν\nu of AA makes the left AA-module isomorphism (2.1) into an AA-bimodule isomorphism

Aν1D(A).{}_{1}A_{\nu}\rightarrow D(A).

Moreover, there exists another AA-bimodule isomorphism

t2:Aν11HomAe(AAe,AeAe)=A;x[yiyuiν(x)vi],t_{2}:{}_{1}A_{\nu^{-1}}\rightarrow\operatorname{Hom}\nolimits_{A^{\textrm{e}}}({}_{A^{\textrm{e}}}A,{}_{A^{\textrm{e}}}A^{\textrm{e}})=A^{\vee};\quad x\mapsto[y\mapsto\sum_{i}yu_{i}\nu(x)\otimes v_{i}],

where the AA-bimodule structure of AA^{\vee} is given by (afb)(x):=f(x)(ba)(afb)(x):=f(x)(b\otimes a^{\circ}) for fAf\in A^{\vee} and a,bAa,b\in A.

2.3. Complete resolutions and their minimalities

Our aim in this subsection is to recall the definition of complete resolutions and to characterize minimal complete resolutions in terms of minimal projective resolutions and minimal injective resolutions. Let us start with the definition of complete resolutions.

Definition 2.1 ([AM02]).
  1. (1)

    A complete resolution of a finitely generated AA-module MM is a diagram

    𝐓=iTiϑ𝐏𝜀M,\operatorname{\mathbf{T}}\nolimits=\bigoplus_{i\in\operatorname{\mathbb{Z}}\nolimits}T_{i}\xrightarrow{\vartheta}\operatorname{\mathbf{P}}\nolimits\xrightarrow{\varepsilon}M,

    where 𝐓\operatorname{\mathbf{T}}\nolimits is an exact sequence of finitely generated projective AA-modules with Hi(𝐓)=0\operatorname{H}\nolimits^{i}(\operatorname{\mathbf{T}}\nolimits^{\vee})=0 for all ii\in\operatorname{\mathbb{Z}}\nolimits, ε:𝐏M\varepsilon:\operatorname{\mathbf{P}}\nolimits\rightarrow M is a projective resolution and ϑ:𝐓𝐏\vartheta:\operatorname{\mathbf{T}}\nolimits\rightarrow\operatorname{\mathbf{P}}\nolimits is a chain map such that ϑi\vartheta_{i} is an isomorphism for i0i\gg 0.

  2. (2)

    A finitely generated AA-module GG is totally reflexive if the canonical morphism G(G)G\rightarrow(G^{\vee})^{\vee} is an isomorphism and ExtAn(G,A)=0=ExtAon(G,A)\operatorname{Ext}\nolimits_{A}^{n}(G,A)=0=\operatorname{Ext}\nolimits_{A^{\textrm{o}}}^{n}(G^{\vee},A) for all n1n\geq 1.

  3. (3)

    A 𝒢\mathcal{G}-resolution ((of length l)\leq l) of a finitely generated AA-module MM is a quasi-isomorphism 𝐆=i0GiM\mathbf{G}=\bigoplus_{i\geq 0}G_{i}\rightarrow M with GiG_{i} totally reflexive ((and Gi=0G_{i}=0 for i>l)i>l).

We define the G-dimension G\mathrm{G}-dimAM\dim_{A}M of a finitely generated AA-module MM by

G-dimAM:=inf{g|there exists a 𝒢-resolution of lengthg}.\mathrm{G}\text{-}\dim_{A}M:=\inf\left\{g\in\operatorname{\mathbb{N}}\nolimits\big{|}\text{there exists a }\mathcal{G}\text{-resolution of length}\leq g\right\}.

Since we are interested in self-injective algebras including Frobenius algebras, we mainly deal with self-injective algebras. For more general cases, we refer to [AM02, BJ13].

Let AA be a self-injective algebra. Since AA is injective as a left and as a right AA-module, any finitely generated AA-module MM is totally reflexive and hence of G-dimension 0. It follows from [AM02, Theorem 3.1] that MM has a complete resolution 𝐓ϑ𝐏M\operatorname{\mathbf{T}}\nolimits\xrightarrow{\vartheta}\operatorname{\mathbf{P}}\nolimits\rightarrow M with ϑi\vartheta_{i} isomorphic for i0i\geq 0. Thus, any complete resolution 𝐓\operatorname{\mathbf{T}}\nolimits of MM consists of some projective resolution of MM in non-negative degrees, and we have M=Cokerd1𝐓M=\operatorname{Coker}\nolimits d^{\operatorname{\mathbf{T}}\nolimits}_{1}. Thus we simply write 𝐓M\operatorname{\mathbf{T}}\nolimits\rightarrow M for 𝐓𝐏M\operatorname{\mathbf{T}}\nolimits\rightarrow\operatorname{\mathbf{P}}\nolimits\rightarrow M.

For a complete resolution 𝐓\operatorname{\mathbf{T}}\nolimits of MM and ii\in\operatorname{\mathbb{Z}}\nolimits, let di𝐓=ιiπid_{i}^{\operatorname{\mathbf{T}}\nolimits}=\iota_{i}\pi_{i} be the canonical factorization through Imdi𝐓\operatorname{Im}\nolimits d_{i}^{\operatorname{\mathbf{T}}\nolimits}, i.e., πi\pi_{i} is the canonical epimorphism TiImdi𝐓T_{i}\rightarrow\operatorname{Im}\nolimits d_{i}^{\operatorname{\mathbf{T}}\nolimits} and ιi\iota_{i} is the canonical inclusion Imdi𝐓Ti1\operatorname{Im}\nolimits d_{i}^{\operatorname{\mathbf{T}}\nolimits}\hookrightarrow T_{i-1}. In particular, we denote by ε\varepsilon the epimorphism π0:T0Imd0𝐓=M\pi_{0}:T_{0}\rightarrow\operatorname{Im}\nolimits d_{0}^{\operatorname{\mathbf{T}}\nolimits}=M and by η\eta the canonical inclusion ι0:MT1\iota_{0}:M\hookrightarrow T_{-1}. The morphism ε:T0M\varepsilon:T_{0}\rightarrow M is called the augmentation of 𝐓\operatorname{\mathbf{T}}\nolimits. Note that the augmentation of any complete resolution of MM induce a chain map 𝐓M\operatorname{\mathbf{T}}\nolimits\rightarrow M, but it is not a quasi-isomorphism. A chain map f:𝐓𝐓f:\operatorname{\mathbf{T}}\nolimits\rightarrow\operatorname{\mathbf{T}}\nolimits^{\prime} between two complete resolutions 𝐓𝜀M\operatorname{\mathbf{T}}\nolimits\xrightarrow{\varepsilon}M and 𝐓εM\operatorname{\mathbf{T}}\nolimits^{\prime}\xrightarrow{\varepsilon^{\prime}}M is called augmentation-preserving if εf=ε\varepsilon^{\prime}f=\varepsilon. It follows from [AM02, Lemma 5.3] that any augmentation-preserving chain map f:𝐓𝐓f:\operatorname{\mathbf{T}}\nolimits\rightarrow\operatorname{\mathbf{T}}\nolimits^{\prime} between two complete resolutions 𝐓\operatorname{\mathbf{T}}\nolimits and 𝐓\operatorname{\mathbf{T}}\nolimits^{\prime} of MM is a homotopy equivalence.

In [AM02], Avramov and Martsinkovsky introduced the notion of minimal complexes. Recall that a chain complex CC over AA is minimal if every homotopy equivalence CCC\rightarrow C is an isomorphism. Clearly, the minimality of a complex preserves under taking shifts. We will apply the notion to complete resolutions over a self-injective algebra.

Definition 2.2.

Let AA be a self-injective algebra and MM a finitely generated AA-module. Then a complete resolution 𝐓𝐏M\operatorname{\mathbf{T}}\nolimits\rightarrow\operatorname{\mathbf{P}}\nolimits\rightarrow M is called minimal if 𝐓\operatorname{\mathbf{T}}\nolimits is minimal.

Remark that our definition of minimal complete resolutions does not require the minimalities of the projective resolutions in non-negative degrees. We now characterize minimal complete resolutions of MM in terms of its projective and injective resolutions. For this purpose, we first recall the result of Avramov and Martsinkovsky. Recall that a projective resolution 𝐏M\operatorname{\mathbf{P}}\nolimits\rightarrow M is a minimal projective resolution if PnP_{n} is a projective cover of Cokerdn+1𝐏\operatorname{Coker}\nolimits d_{n+1}^{\operatorname{\mathbf{P}}\nolimits} for all n0n\geq 0 and that an injective resolution M𝐈=i0IiM\rightarrow\operatorname{\mathbf{I}}\nolimits=\bigoplus_{i\leq 0}I_{i} is a minimal injective resolution if InI_{n} is an injective envelope of Kerdn𝐈\operatorname{Ker}\nolimits d_{n}^{\operatorname{\mathbf{I}}\nolimits} for all n0n\leq 0.

Lemma 2.3 ([AM02, Example 1.8]).

Let MM be a finitely generated AA-module, and let 𝐏M\operatorname{\mathbf{P}}\nolimits\rightarrow M be a projective resolution and M𝐈M\rightarrow\operatorname{\mathbf{I}}\nolimits an injective resolution. Then the following statements hold.

  1. (1)

    𝐏\operatorname{\mathbf{P}}\nolimits is minimal if and only if PnP_{n} is a projective cover of Cokerdn+1𝐏\operatorname{Coker}\nolimits d_{n+1}^{\operatorname{\mathbf{P}}\nolimits} for all n0n\geq 0.

  2. (2)

    𝐈\operatorname{\mathbf{I}}\nolimits is minimal if and only if InI_{n} is an injective envelope of Kerdn𝐈\operatorname{Ker}\nolimits d_{n}^{\operatorname{\mathbf{I}}\nolimits} for all n0n\leq 0.

Proof.

we only prove (1)(1); the proof of (2)(2) is similar. It follows from [AM02, Proposition 1.7(1)] that 𝐏\operatorname{\mathbf{P}}\nolimits is minimal if and only if each chain map f:𝐏𝐏f:\operatorname{\mathbf{P}}\nolimits\rightarrow\operatorname{\mathbf{P}}\nolimits homotopic to id𝐏\operatorname{id}\nolimits_{\operatorname{\mathbf{P}}\nolimits} is an isomorphism. Take a chain map f:𝐏𝐏f:\operatorname{\mathbf{P}}\nolimits\rightarrow\operatorname{\mathbf{P}}\nolimits such that fid𝐏f\sim\operatorname{id}\nolimits_{\operatorname{\mathbf{P}}\nolimits}. Then there exists a morphism h0:P0P1h_{0}:P_{0}\rightarrow P_{1} such that idP0f0=d1h0\operatorname{id}\nolimits_{P_{0}}-f_{0}=d_{1}h_{0}. Letting ε\varepsilon be the augmentation P0MP_{0}\rightarrow M, we have ε(idf0)=ε(d1h0)=0\varepsilon(\operatorname{id}\nolimits-f_{0})=\varepsilon(d_{1}h_{0})=0. Since ε\varepsilon is a projective cover of MM, the morphism f0f_{0} is an epimorphism and hence an isomorphism. Moreover, it induces a commutative square

P1\textstyle{P_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f1\scriptstyle{f_{1}}Cokerd2𝐏\textstyle{\operatorname{Coker}\nolimits d^{\operatorname{\mathbf{P}}\nolimits}_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f1¯\scriptstyle{\overline{f_{1}}}P1\textstyle{P_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Cokerd2𝐏\textstyle{\operatorname{Coker}\nolimits d^{\operatorname{\mathbf{P}}\nolimits}_{2}}

where the morphism f1¯\overline{f_{1}} induced by f1f_{1} is an isomorphism and the horizontal morphisms are the canonical epimorphisms. Since the canonical epimorphism P1Cokerd2𝐏P_{1}\rightarrow\operatorname{Coker}\nolimits d^{\operatorname{\mathbf{P}}\nolimits}_{2} is a projective cover, the morphism f1f_{1} is an isomorphism. Inductively, we see that the morphism fif_{i} is an isomorphism for all i0i\geq 0.

Conversely, suppose that 𝐏\operatorname{\mathbf{P}}\nolimits is minimal. Observe that any projective resolution 𝐏\operatorname{\mathbf{P}}\nolimits of MM can be decomposed as 𝐏=𝐏M𝐏\operatorname{\mathbf{P}}\nolimits=\operatorname{\mathbf{P}}\nolimits_{M}\oplus\operatorname{\mathbf{P}}\nolimits^{\prime}, where 𝐏M\operatorname{\mathbf{P}}\nolimits_{M} is a minimal projective resolution of MM and 𝐏\operatorname{\mathbf{P}}\nolimits^{\prime} is a contractible complex. It follows from [AM02, Proposition 1.7(3)] that 𝐏\operatorname{\mathbf{P}}\nolimits^{\prime} must be zero. This completes the proof. ∎

Let AA be a self-injective algebra and 𝐓\operatorname{\mathbf{T}}\nolimits a complete resolution of a finitely generated AA-module MM with d0𝐓=ηεd_{0}^{\operatorname{\mathbf{T}}\nolimits}=\eta\varepsilon. Let 𝐓0:=i0Ti\operatorname{\mathbf{T}}\nolimits_{\geq 0}:=\bigoplus_{i\geq 0}T_{i} and 𝐓<0:=i<0Ti\operatorname{\mathbf{T}}\nolimits_{<0}:=\bigoplus_{i<0}T_{i} be the truncated subcomplexes of 𝐓\operatorname{\mathbf{T}}\nolimits with the inherited differentials, which are of the forms

𝐓0:\textstyle{\operatorname{\mathbf{T}}\nolimits_{\geq 0}:\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}T2\textstyle{T_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d2\scriptstyle{d_{2}}T1\textstyle{T_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d1\scriptstyle{d_{1}}T0\textstyle{T_{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces},\textstyle{\cdots,}𝐓<0:\textstyle{\operatorname{\mathbf{T}}\nolimits_{<0}:\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}T1\textstyle{T_{-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d1\scriptstyle{d_{-1}}T2\textstyle{T_{-2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d2\scriptstyle{d_{-2}}T3\textstyle{T_{-3}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}.\textstyle{\cdots.}

Note that 𝐓0\operatorname{\mathbf{T}}\nolimits_{\geq 0} is a projective resolution of MM and that 𝐓<0\operatorname{\mathbf{T}}\nolimits_{<0} is isomorphic to 𝐈[1]\operatorname{\mathbf{I}}\nolimits[-1] for some injective resolution 𝐈\operatorname{\mathbf{I}}\nolimits of MM.

Proposition 2.4.

Under the same notation above, the following statements are equivalent.

  1. (1)

    𝐓\operatorname{\mathbf{T}}\nolimits is a minimal complete resolution of MM.

  2. (2)

    TnT_{n} is a projective cover of Cokerdn+1𝐓\operatorname{Coker}\nolimits d_{n+1}^{\operatorname{\mathbf{T}}\nolimits} for all nn\in\operatorname{\mathbb{Z}}\nolimits.

Proof.

Assume that 𝐓\operatorname{\mathbf{T}}\nolimits is minimal. Since the (n)(-n)-shifted complex 𝐓[n]\operatorname{\mathbf{T}}\nolimits[-n] is a minimal complete resolution of ΩAn(M)\Omega_{A}^{n}(M) for any nn\in\operatorname{\mathbb{Z}}\nolimits, it suffices to show that the augmentation ε:P0M\varepsilon:P_{0}\rightarrow M is a projective cover. Let NN be an AA-module and f:NP0f:N\rightarrow P_{0} be a morphism such that the composite εf\varepsilon f is an epimorphism. The projectivity of P0P_{0} implies that there exists a morphism g:P0Ng:P_{0}\rightarrow N such that ε=ε(fg)\varepsilon=\varepsilon(fg). By Lemmas 1.1 and 1.2, the morphism fgfg extends to a chain map φ:𝐓𝐓\varphi:\operatorname{\mathbf{T}}\nolimits\rightarrow\operatorname{\mathbf{T}}\nolimits satisfying εφ0=εfg=ε\varepsilon\varphi_{0}=\varepsilon fg=\varepsilon. It follows from [AM02, Lemma 5.3] that the chain map φ\varphi is homotopy equivalent. The minimality of 𝐓\operatorname{\mathbf{T}}\nolimits implies that φ0=fg\varphi_{0}=fg is an isomorphism. Therefore, ff is an epimorphism.

Conversely, thanks to [AM02, Proposition 1.7(1)], it suffices to prove the converse for a chain map f:𝐓𝐓f:\operatorname{\mathbf{T}}\nolimits\rightarrow\operatorname{\mathbf{T}}\nolimits such that fid𝐓f\sim\operatorname{id}\nolimits_{\operatorname{\mathbf{T}}\nolimits}. Take a homotopy hh from id𝐓\operatorname{id}\nolimits_{\operatorname{\mathbf{T}}\nolimits} to ff and define a graded map φ=(φi)i:𝐓𝐓\varphi=(\varphi_{i})_{i\in\operatorname{\mathbb{Z}}\nolimits}:\operatorname{\mathbf{T}}\nolimits\rightarrow\operatorname{\mathbf{T}}\nolimits by

φi={fi if i0,1,f0+h1d0𝐓 if i=0,f1+d0𝐓h1 if i=1.\varphi_{i}=\begin{cases}f_{i}&\mbox{ if $i\not=0,-1$,}\\[3.0pt] f_{0}+h_{-1}d_{0}^{\operatorname{\mathbf{T}}\nolimits}&\mbox{ if $i=0$,}\\[3.0pt] f_{-1}+d_{0}^{\operatorname{\mathbf{T}}\nolimits}h_{-1}&\mbox{ if $i=-1$}.\end{cases}

Then φ\varphi is a chain map such that εφ0=ε\varepsilon\varphi_{0}=\varepsilon. Note that 𝐓0\operatorname{\mathbf{T}}\nolimits_{\geq 0} is a minimal projective resolution of MM. Since the chain map φ0:𝐓0𝐓0\varphi_{\geq 0}:\operatorname{\mathbf{T}}\nolimits_{\geq 0}\rightarrow\operatorname{\mathbf{T}}\nolimits_{\geq 0} is homotopy equivalent, it follows from Lemma 2.3(1) that each φi=fi\varphi_{i}=f_{i} with i>0i>0 is an isomorphism. Since the inclusion ιn\iota_{n} with n0n\leq 0 is an injective envelope of Kerdn1𝐓\operatorname{Ker}\nolimits d_{n-1}^{\operatorname{\mathbf{T}}\nolimits}, Lemma 2.3(2) and the fact that φ1η=η\varphi_{-1}\eta=\eta yield that each φi=fi\varphi_{i}=f_{i} with i<1i<-1 is an isomorphism. It remains to show that f0f_{0} and f1f_{-1} are isomorphisms. Since f1f_{1} is an isomorphism, so is the restriction f0~\widetilde{f_{0}} of f0f_{0} to Kerd0𝐓\operatorname{Ker}\nolimits d_{0}^{\operatorname{\mathbf{T}}\nolimits}. In a commutative square

Kerd0𝐓\textstyle{\operatorname{Ker}\nolimits d_{0}^{\operatorname{\mathbf{T}}\nolimits}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ι1\scriptstyle{\iota_{1}}f0~\scriptstyle{\widetilde{f_{0}}}T0\textstyle{T_{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f0\scriptstyle{f_{0}}Kerd0𝐓\textstyle{\operatorname{Ker}\nolimits d_{0}^{\operatorname{\mathbf{T}}\nolimits}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ι1\scriptstyle{\iota_{1}}T0\textstyle{T_{0}}

the fact that ι1\iota_{1} is an injective envelope of Kerd0𝐓\operatorname{Ker}\nolimits d_{0}^{\operatorname{\mathbf{T}}\nolimits} implies that f0f_{0} is a monomorphism and thus an isomorphism. Similarly, one shows that f1f_{-1} is an isomorphism. ∎

In the course of the proof of Proposition 2.4, we have proved the following.

Corollary 2.5.

Let AA be a self-injective algebra and MM a finitely generated AA-module. Then any minimal complete resolution of MM is isomorphic to the complete resolution of the form

𝐏\textstyle{\operatorname{\mathbf{P}}\nolimits\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ε\scriptstyle{\varepsilon}\scriptstyle{\circlearrowright}𝚺1𝐈\textstyle{\boldsymbol{\Sigma}^{-1}\operatorname{\mathbf{I}}\nolimits}M\textstyle{M\ignorespaces\ignorespaces\ignorespaces\ignorespaces}η\scriptstyle{-\eta}

where 𝐏𝜀M\operatorname{\mathbf{P}}\nolimits\xrightarrow{\varepsilon}M is a minimal projective resolution and M𝜂𝐈M\xrightarrow{\eta}\operatorname{\mathbf{I}}\nolimits is a minimal injective resolution.

It follows from Corollary 2.5 that any finitely generated AA-module admits a minimal complete resolution. It follows from [AM02, Proposition 1.7(2)] that a minimal complete resolution is uniquely determined up to isomorphism. For a minimal complete resolution 𝐓\operatorname{\mathbf{T}}\nolimits of a finitely generated AA-module MM and ii\in\operatorname{\mathbb{Z}}\nolimits, we set ΩAiM:=Kerdi1𝐓\Omega_{A}^{i}M:=\operatorname{Ker}\nolimits d_{i-1}^{\operatorname{\mathbf{T}}\nolimits}. Note that Corollary 2.5 implies that the module ΩAiM\Omega_{A}^{i}M is nothing but the syzygy module of MM if i0i\geq 0 and the cosyzygy module of MM if i1i\leq-1 (see [SY11] for (co)syzygy modules).

2.4. Tate and Tate-Hochschild (co)homology groups

In this subsection, we recall the definition of Tate and Tate-Hochschild (co)homology groups and show that there exists certain duality between Tate-Hochschild cohomology and homology groups. Recall that if AA is a self-injective (Frobenius) algebra, then so is the enveloping algebra AeA^{\textrm{e}}.

Definition 2.6.

Let AA be a self-injective algebra and nn\in\operatorname{\mathbb{Z}}\nolimits.

  1. (1)

    Let MM be a finitely generated left AA-module with a complete resolution 𝐓M\operatorname{\mathbf{T}}\nolimits^{M}, LL a finitely generated right AA-module with a complete resolution 𝐓L\operatorname{\mathbf{T}}\nolimits^{L} and NN a left AA-module. The nn-th Tate cohomology group Ext^An(M,N)\operatorname{\widehat{Ext}}\nolimits_{A}^{n}(M,N) is defined by

    Ext^An(M,N):=Hn(𝓂A(𝐓M,N)),\operatorname{\widehat{Ext}}\nolimits_{A}^{n}(M,N):=\operatorname{H}\nolimits^{n}(\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A}(\operatorname{\mathbf{T}}\nolimits^{M},N)),

    and the nn-th Tate homology group Tor^nA(M,N)\operatorname{\widehat{Tor}}\nolimits^{A}_{n}(M,N) is defined by

    Tor^nA(L,N):=Hn(𝐓LAN).\operatorname{\widehat{Tor}}\nolimits_{n}^{A}(L,N):=\operatorname{H}\nolimits_{n}(\operatorname{\mathbf{T}}\nolimits^{L}\otimes_{A}N).
  2. (2)

    The nn-th Tate-Hochschild cohomology group H^n(A,N)\operatorname{\widehat{H}}\nolimits^{n}(A,N) of AA with coefficients in an AA-bimodule NN is defined by

    H^n(A,N):=Ext^Aen(A,N).\operatorname{\widehat{H}}\nolimits^{n}(A,N):=\operatorname{\widehat{Ext}}\nolimits_{A^{\textrm{e}}}^{n}(A,N).

    The nn-th Tate-Hochschild homology group H^n(A,N)\operatorname{\widehat{H}}\nolimits_{n}(A,N) of AA with coefficients in NN by

    H^n(A,N):=Tor^nAe(A,N).\operatorname{\widehat{H}}\nolimits_{n}(A,N):=\operatorname{\widehat{Tor}}\nolimits_{n}^{A^{\textrm{e}}}(A,N).

Let us recall the definitions of projectively stable categories and of stable tensor products: the projectively stable category AA-mod¯\underline{\mathrm{mod}} of AA is the category whose objects are finitely generated left AA-modules and whose morphisms are given by the quotient space

Hom¯A(M,N):=HomA(M,N)/𝒫(M,N),\displaystyle\underline{\operatorname{Hom}\nolimits}_{A}(M,N):=\operatorname{Hom}\nolimits_{A}(M,N)/\mathcal{P}(M,N),

where 𝒫(M,N)\mathcal{P}(M,N) is the space of morphisms factoring through a projective AA-module. The stable tensor product L¯ANL\underline{\otimes}_{A}N of a finitely generated right AA-module LL with a finitely generated left AA-module NN is defined to be

L¯AN:={wLAN|wι:NII:injectiveKer(idLAι)}.\displaystyle L\underline{\otimes}_{A}N:=\Bigg{\{}w\in L\otimes_{A}N\ \Bigg{|}\ w\in\bigcap_{\begin{subarray}{c}\iota:N\hookrightarrow I\\ I:{\footnotesize\mbox{injective}}\end{subarray}}\operatorname{Ker}\nolimits(\operatorname{id}\nolimits_{L}\otimes_{A}\iota)\Bigg{\}}.

In case of Frobenius algebras, it follows from [ES09, Proposition 2.1.3] that there exists an isomorphism

L¯AN{wLAN|wι:LII:injectiveKer(ιAidN)}.\displaystyle L\underline{\otimes}_{A}N\cong\Bigg{\{}w\in L\otimes_{A}N\ \Bigg{|}\ w\in\bigcap_{\begin{subarray}{c}\iota:L\hookrightarrow I\\ I:{\footnotesize\mbox{injective}}\end{subarray}}\operatorname{Ker}\nolimits(\iota\otimes_{A}\operatorname{id}\nolimits_{N})\Bigg{\}}.

We will identify the two modules above via this isomorphism. Remark that [ES09, Proposition 2.1.3] also holds for self-injective algebras, because the key of the proof is the projectivity of an injective module.

It is well-known that there exist isomorphisms

ExtAi(M,N)Hom¯A(ΩAiM,N),ToriA(L,N)ΩAiL¯AN\displaystyle\operatorname{Ext}\nolimits_{A}^{i}(M,N)\cong\underline{\operatorname{Hom}\nolimits}_{A}(\Omega_{A}^{i}M,N),\quad\operatorname{Tor}\nolimits_{i}^{A}(L,N)\cong\Omega_{A}^{i}L\underline{\otimes}_{A}N

for i1i\geq 1 when AA is a self-injective algebra. It is known that there exist such isomorphisms even for Tate (co)homology groups. We will include a proof.

Proposition 2.7.

Let AA be a self-injective algebra, LL a finitely generated right AA-module and MM and NN finitely generated left AA-modules. Then there exist isomorphisms

Ext^Ai(M,N)Hom¯A(ΩAiM,N),Tor^iA(L,N)ΩAiL¯AN\displaystyle\operatorname{\widehat{Ext}}\nolimits_{A}^{i}(M,N)\cong\underline{\operatorname{Hom}\nolimits}_{A}(\Omega_{A}^{i}M,N),\quad\operatorname{\widehat{Tor}}\nolimits_{i}^{A}(L,N)\cong\Omega_{A}^{i}L\underline{\otimes}_{A}N

for ii\in\operatorname{\mathbb{Z}}\nolimits.

Proof.

Let ii\in\operatorname{\mathbb{Z}}\nolimits be fixed. For the first isomorphism, let 𝐓\operatorname{\mathbf{T}}\nolimits be a minimal complete resolution of MM. Take fHomA(ΩAiM,N)f\in\operatorname{Hom}\nolimits_{A}(\Omega_{A}^{i}M,N) and consider a commutative diagram

Ti\textstyle{T_{i}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}di𝐓\scriptstyle{d_{i}^{\operatorname{\mathbf{T}}\nolimits}}πi\scriptstyle{\pi_{i}}fπi\scriptstyle{f\pi_{i}}Ti1\textstyle{T_{i-1}}ΩAiM\textstyle{\Omega_{A}^{i}M\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ιi\scriptstyle{\iota_{i}}f\scriptstyle{f}N\textstyle{N}

Obviously, fπif\pi_{i} belongs to Kerdi𝓂A(𝐓,N)\operatorname{Ker}\nolimits d^{\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A}(\operatorname{\mathbf{T}}\nolimits,N)}_{i}. If ff factors through a projective AA-module PP, then fπi𝓂A(𝐓,N)if\pi_{i}\in\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A}(\operatorname{\mathbf{T}}\nolimits,N)^{i} is a coboundary, because the projective module PP is injective. Thus we have a well-defined morphism

Φi:Hom¯A(ΩAiM,N)Ext^Ai(M,N)\displaystyle\Phi_{i}:\underline{\operatorname{Hom}\nolimits}_{A}(\Omega_{A}^{i}M,N)\rightarrow\operatorname{\widehat{Ext}}\nolimits_{A}^{i}(M,N)

given by [f][fπi][f]\mapsto[f\pi_{i}]. We claim that Φi\Phi_{i} is an isomorphism. If g𝓂A(𝐓,N)ig\in\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A}(\operatorname{\mathbf{T}}\nolimits,N)^{i} lies in Kerdi𝓂A(𝐓,N)\operatorname{Ker}\nolimits d^{\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A}(\operatorname{\mathbf{T}}\nolimits,N)}_{i}, then there uniquely exists a morphism g:ΩAiMNg^{\prime}:\Omega_{A}^{i}M\rightarrow N such that g=gπig=g^{\prime}\pi_{i}. This implies that Φi\Phi_{i} is an epimorphism. Assume now that fπif\pi_{i} is a coboundary, that is, fπi=(1)ifdi𝐓f\pi_{i}=(-1)^{i}f^{\prime}d^{\operatorname{\mathbf{T}}\nolimits}_{i} for some f:Ti1Nf^{\prime}:T_{i-1}\rightarrow N. The surjectivity of πi\pi_{i} yields that (1)ifι=f(-1)^{i}f^{\prime}\iota=f, which means that Φi\Phi_{i} is a monomorphism.

For the second isomorphism, let 𝐓\operatorname{\mathbf{T}}\nolimits be a minimal complete resolution of LL. Consider a commutative diagram

Ti+1AN\textstyle{T_{i+1}\otimes_{A}N\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}di+1𝐓Aid\scriptstyle{d_{i+1}^{\operatorname{\mathbf{T}}\nolimits}\otimes_{A}\operatorname{id}\nolimits}πi+1Aid\scriptstyle{\pi_{i+1}\otimes_{A}\operatorname{id}\nolimits}TiAN\textstyle{T_{i}\otimes_{A}N\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}di𝐓Aid\scriptstyle{d_{i}^{\operatorname{\mathbf{T}}\nolimits}\otimes_{A}\operatorname{id}\nolimits}πiAid\scriptstyle{\pi_{i}\otimes_{A}\operatorname{id}\nolimits}Ti1AN\textstyle{T_{i-1}\otimes_{A}N}ΩAi+1LAN\textstyle{\Omega_{A}^{i+1}L\otimes_{A}N\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ιi+1Aid\scriptstyle{\iota_{i+1}\otimes_{A}\operatorname{id}\nolimits}ΩAiLAN\textstyle{\Omega_{A}^{i}L\otimes_{A}N\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ιiAid\scriptstyle{\iota_{i}\otimes_{A}\operatorname{id}\nolimits}

Let wΩAiL¯ANw\in\Omega_{A}^{i}L\underline{\otimes}_{A}N be arbitrary. Since πiAeidN\pi_{i}\otimes_{A^{\textrm{e}}}\operatorname{id}\nolimits_{N} is an epimorphism, there exists zTiANz\in T_{i}\otimes_{A}N such that w=(πiAeidN)(z)w=(\pi_{i}\otimes_{A^{\textrm{e}}}\operatorname{id}\nolimits_{N})(z). By the definition of ΩAiL¯AN\Omega_{A}^{i}L\underline{\otimes}_{A}N and the injectivity of Ti1T_{i-1}, we have (di𝐓AidN)(z)=(ιiAidN)(w)=0(d_{i}^{\operatorname{\mathbf{T}}\nolimits}\otimes_{A}\operatorname{id}\nolimits_{N})(z)=(\iota_{i}\otimes_{A}\operatorname{id}\nolimits_{N})(w)=0. If we take zTiANz^{\prime}\in T_{i}\otimes_{A}N such that w=(πiAeidN)(z)w=(\pi_{i}\otimes_{A^{\textrm{e}}}\operatorname{id}\nolimits_{N})(z^{\prime}), then zzz-z^{\prime} belongs to Im(di+1𝐓AidN)\operatorname{Im}\nolimits(d_{i+1}^{\operatorname{\mathbf{T}}\nolimits}\otimes_{A}\operatorname{id}\nolimits_{N}) because zzKer(πiAidN)z-z^{\prime}\in\operatorname{Ker}\nolimits(\pi_{i}\otimes_{A}\operatorname{id}\nolimits_{N}). Thus we have a well-defined morphism

Ψi:ΩAiL¯ANTor^iA(L,N)\displaystyle\Psi_{i}:\Omega_{A}^{i}L\underline{\otimes}_{A}N\rightarrow\operatorname{\widehat{Tor}}\nolimits_{i}^{A}(L,N)

given by w[z]w\mapsto[z]. It is easy to check that Ψi\Psi_{i} is an isomorphism. ∎

Remark 2.8.

From Proposition 2.7, we have isomorphisms

H^i(A,M)Hom¯Ae(ΩAeiA,M)andH^i(A,M)ΩAeiA¯AeM.\displaystyle\operatorname{\widehat{H}}\nolimits^{i}(A,M)\cong\underline{\operatorname{Hom}\nolimits}_{A^{\textrm{e}}}(\Omega_{A^{\textrm{e}}}^{i}A,M)\ \ \mbox{and}\ \ \operatorname{\widehat{H}}\nolimits_{i}(A,M)\cong\Omega_{A^{\textrm{e}}}^{i}A\underline{\otimes}_{A^{\textrm{e}}}M.

The vector spaces Hom¯Ae(ΩAeiA,M)\underline{\operatorname{Hom}\nolimits}_{A^{\textrm{e}}}(\Omega_{A^{\textrm{e}}}^{i}A,M) and ΩAeiA¯AeM\Omega_{A^{\textrm{e}}}^{i}A\underline{\otimes}_{A^{\textrm{e}}}M are known to be the stable Hochschild cohomology and homology groups defined by Eu and Shedler (see [ES09]).

The following lemma is the dual of [BJ13, Lemma 3.6].

Lemma 2.9.

Let MM and NN be AA-modules and αAut(A)\alpha\in\operatorname{Aut}\nolimits(A). Then there exists an isomorphism

ExtAi(Mα,N)ExtAi(M,Nα1)\operatorname{Ext}\nolimits_{A}^{i}(\,{}_{\alpha}M,N)\cong\operatorname{Ext}\nolimits_{A}^{i}(M,{}_{\alpha^{-1}}N)

for all i0i\geq 0. Moreover, if AA is a self-injective algebra, and MM is finitely generated, then there exists an isomorphism

Ext^Ai(Mα,N)Ext^Ai(M,Nα1)\operatorname{\widehat{Ext}}\nolimits_{A}^{i}(\,{}_{\alpha}M,N)\cong\operatorname{\widehat{Ext}}\nolimits_{A}^{i}(M,{}_{\alpha^{-1}}N)

for all ii\in\operatorname{\mathbb{Z}}\nolimits.

Proof.

Let 𝐏\operatorname{\mathbf{P}}\nolimits be a projective resolution of MM. It follows from the proof of [BJ13, Lemma 3.6] that 𝐏α{}_{\alpha}\operatorname{\mathbf{P}}\nolimits is a projective resolution of Mα{}_{\alpha}M. Thus we have isomorphisms

ExtAi(Mα,N)\displaystyle\operatorname{Ext}\nolimits_{A}^{i}(\,{}_{\alpha}M,N) Hi(𝓂A(α𝐏,N))\displaystyle\cong\operatorname{H}\nolimits^{i}(\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A}(_{\alpha}\operatorname{\mathbf{P}}\nolimits,N))
Hi(𝓂A(𝐏,Nα1))ExtAi(M,Nα1).\displaystyle\cong\operatorname{H}\nolimits^{i}(\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A}(\operatorname{\mathbf{P}}\nolimits,{}_{\alpha^{-1}}N))\cong\operatorname{Ext}\nolimits_{A}^{i}(M,{}_{\alpha^{-1}}N).

Assume now that AA is a self-injective algebra and that MM is finitely generated. The proof of [BJ13, Lemma 3.6] implies that if 𝐓\operatorname{\mathbf{T}}\nolimits is a complete resolution of MM, then 𝐓α{}_{\alpha}\operatorname{\mathbf{T}}\nolimits is a complete resolution of Mα{}_{\alpha}M. Replacing 𝐏\operatorname{\mathbf{P}}\nolimits by 𝐓\operatorname{\mathbf{T}}\nolimits in the argument above yields the later statement. ∎

For any automorphisms α,βAut(A)\alpha,\beta\in\operatorname{Aut}\nolimits(A), the twisted complex Cβα{}_{\alpha}C_{\beta} is defined to be the chain complex with components (Cβα)n=(Cn)βα({}_{\alpha}C_{\beta})_{n}={}_{\alpha}(C_{n})_{\beta} and the inherited differentials.

In the rest of this paper, we assume that all complete resolutions of a self-injective algebra AA are taken over AeA^{\textrm{e}}. Using the two truncations 𝐓0\operatorname{\mathbf{T}}\nolimits_{\geq 0} and 𝐓<0\operatorname{\mathbf{T}}\nolimits_{<0}, we can write 𝐓\operatorname{\mathbf{T}}\nolimits as 𝐓0d0𝐓𝐓<0\operatorname{\mathbf{T}}\nolimits_{\geq 0}\xrightarrow{d_{0}^{\operatorname{\mathbf{T}}\nolimits}}\operatorname{\mathbf{T}}\nolimits_{<0} with d0𝐓=ηεd_{0}^{\operatorname{\mathbf{T}}\nolimits}=\eta\varepsilon. The quasi-isomorphism A𝜂𝐓<0A\xrightarrow{\eta}\operatorname{\mathbf{T}}\nolimits_{<0} is called a backward projective resolution of AA.

Lemma 2.10.

Let AA be a Frobenius algebra with the Nakayama automorphism ν\nu and 𝐓\operatorname{\mathbf{T}}\nolimits arbitrary complete resolution of AA. Then there exist two projective resolutions 𝐏𝜀A\operatorname{\mathbf{P}}\nolimits\xrightarrow{\varepsilon}A and 𝐏εA\operatorname{\mathbf{P}}\nolimits^{\prime}\xrightarrow{\varepsilon^{\prime}}A such that 𝐓\operatorname{\mathbf{T}}\nolimits is isomorphic to 𝐏d0((𝚺𝐏))ν1\operatorname{\mathbf{P}}\nolimits\xrightarrow{d_{0}}{}_{1}((\boldsymbol{\Sigma}\operatorname{\mathbf{P}}\nolimits^{\prime})^{\vee})_{\nu}, where d0d_{0} is the composition of ε\varepsilon with η:=(ε)t2\eta^{\prime}:=(\varepsilon^{\prime})^{\vee}t_{2}.

Proof.

It suffices to show that the backward projective resolution A𝜂𝐓<0A\xrightarrow{\eta}\operatorname{\mathbf{T}}\nolimits_{<0} is obtained from some projective resolution in the desired way.

Take the AeA^{\textrm{e}}-dual complex (𝚺𝐓<0)(\boldsymbol{\Sigma}\operatorname{\mathbf{T}}\nolimits_{<0})^{\vee} of the acyclic complex 𝚺𝐓<0\boldsymbol{\Sigma}\operatorname{\mathbf{T}}\nolimits_{<0}. Twisting it by ν1\nu^{-1} on the left hand side, we have a projective resolution

𝐏:=((𝚺𝐓<0))1ν1η(A)1ν1A.\operatorname{\mathbf{P}}\nolimits^{\prime}:={}_{\nu^{-1}}((\boldsymbol{\Sigma}\operatorname{\mathbf{T}}\nolimits_{<0})^{\vee})_{1}\xrightarrow{-\eta^{\vee}}{}_{\nu^{-1}}(A^{\vee})_{1}\cong A.

Note that (Ti)1ν1{}_{\nu^{-1}}(T_{i}^{\vee})_{1} with i1i\leq-1 is a finitely generated projective AA-bimodule. Again, take the AeA^{\textrm{e}}-dual complex (𝚺𝐏)(\boldsymbol{\Sigma}\operatorname{\mathbf{P}}\nolimits^{\prime})^{\vee} of 𝚺𝐏\boldsymbol{\Sigma}\operatorname{\mathbf{P}}\nolimits^{\prime}. Twisting it by ν\nu on the right hand side, we get a backward projective resolution

At2(A)ν1η((𝚺𝐏))ν1=((((𝐓<0))1ν1))ν1.\displaystyle A\xrightarrow{t_{2}}{}_{1}(A^{\vee})_{\nu}\xrightarrow{\eta^{\vee\vee}}{}_{1}((\boldsymbol{\Sigma}\operatorname{\mathbf{P}}\nolimits^{\prime})^{\vee})_{\nu}={}_{1}(({}_{\nu^{-1}}((\operatorname{\mathbf{T}}\nolimits_{<0})^{\vee})_{1})^{\vee})_{\nu}.

This is isomorphic to A𝜂𝐓<0A\xrightarrow{\eta}\operatorname{\mathbf{T}}\nolimits_{<0}. Indeed, there are isomorphisms of complexes

((((𝐓<0))1ν1))ν1=\displaystyle{}_{1}(({}_{\nu^{-1}}((\operatorname{\mathbf{T}}\nolimits_{<0})^{\vee})_{1})^{\vee})_{\nu}= 𝓂Aeidν(𝓂Ae(𝐓<0,Ae)idν1,Ae)\displaystyle\ {}_{\operatorname{id}\nolimits\otimes\nu}\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits_{<0},A^{\textrm{e}})_{\operatorname{id}\nolimits\otimes\nu^{-1}},A^{\textrm{e}})
\displaystyle\cong 𝓂Aeidν(𝓂Ae(𝐓<0,Ae),Aidνe)\displaystyle\ {}_{\operatorname{id}\nolimits\otimes\nu}\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits_{<0},A^{\textrm{e}}),A^{\textrm{e}}_{\operatorname{id}\nolimits\otimes\nu})
\displaystyle\cong 𝓂Aeidν(𝓂Ae(𝐓<0,Ae),Aeidν1)\displaystyle\ {}_{\operatorname{id}\nolimits\otimes\nu}\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits_{<0},A^{\textrm{e}}),{}_{\operatorname{id}\nolimits\otimes\nu^{-1}}A^{\textrm{e}})
=\displaystyle= 𝓂Ae(𝓂Ae(𝐓<0,Ae),Ae)\displaystyle\ \operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits_{<0},A^{\textrm{e}}),A^{\textrm{e}})
\displaystyle\cong 𝐓<0,\displaystyle\ \operatorname{\mathbf{T}}\nolimits_{<0},

where the second isomorphism is induced by AeA^{\textrm{e}}-bimodule isomorphism AidνeAeidν1A^{\textrm{e}}_{\operatorname{id}\nolimits\otimes\nu}\cong{}_{\operatorname{id}\nolimits\otimes\nu^{-1}}A^{\textrm{e}}. ∎

We now recall the description of Tate-Hochschild (co)homology groups: let 𝐓\operatorname{\mathbf{T}}\nolimits be a complete resolution of a Frobenius algebra AA, MM an AA-bimodule, C+:=𝓂Ae(𝐓0,M)C^{+}:=\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits_{\geq 0},M) and C:=𝓂Ae(𝐓<0,M)C^{-}:=\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits_{<0},M). Since 𝐓0\operatorname{\mathbf{T}}\nolimits_{\geq 0} is a projective resolution of AA, we have Hi(C+)=Hi(A,M)\operatorname{H}\nolimits^{i}(C^{+})=\operatorname{H}\nolimits^{i}(A,M) for i0i\geq 0. It follows from Lemma 2.10 that 𝐓<0((𝚺𝐏))ν1\operatorname{\mathbf{T}}\nolimits_{<0}\cong{}_{1}((\boldsymbol{\Sigma}\operatorname{\mathbf{P}}\nolimits)^{\vee})_{\nu} for some projective resolution 𝐏\operatorname{\mathbf{P}}\nolimits of AA. Thus, we have

C\displaystyle C^{-} =𝓂Ae(𝐓<0,M)𝓂Ae(((𝚺𝐏))ν1,M)\displaystyle=\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits_{<0},M)\cong\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}({}_{1}((\boldsymbol{\Sigma}\operatorname{\mathbf{P}}\nolimits)^{\vee})_{\nu},M)
𝓂Ae((𝚺𝐏),Mν11)𝚺𝐏AeMν11\displaystyle\cong\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}((\boldsymbol{\Sigma}\operatorname{\mathbf{P}}\nolimits)^{\vee},{}_{1}M_{\nu^{-1}})\cong\boldsymbol{\Sigma}\operatorname{\mathbf{P}}\nolimits\otimes_{A^{\textrm{e}}}{}_{1}M_{\nu^{-1}}

and hence Hi(C)Hi1(A,Mν11)\operatorname{H}\nolimits^{i}(C^{-})\cong\operatorname{H}\nolimits_{-i-1}(A,{}_{1}M_{\nu^{-1}}) for i1i\leq-1. Therefore, we get

H^n(A,M)={Hn(A,M) if n>0,Hn1(A,Mν11) if n<1.\displaystyle\operatorname{\widehat{H}}\nolimits^{n}(A,M)=\begin{cases}\operatorname{H}\nolimits^{n}(A,M)&\mbox{ if }n>0,\\ \operatorname{H}\nolimits_{-n-1}(A,{}_{1}M_{\nu^{-1}})&\mbox{ if }n<-1.\end{cases}

For the other two cohomology groups, there are isomorphisms

(2.2) H^1(A,M)(MNA)/IA(M),H^0(A,M)MA/NA(M),\displaystyle\operatorname{\widehat{H}}\nolimits^{-1}(A,M)\cong({}_{N_{A}}M)/I_{A}(M),\qquad\operatorname{\widehat{H}}\nolimits^{0}(A,M)\cong M^{A}/N_{A}(M),

where we set

MA:={mM|am=ma for all aA},\displaystyle M^{A}:=\{\,m\in M\ |\ am=ma\mbox{ for all }a\in A\},
NA(M):={iuimvi|mM},MNA:={mM|iuimvi=0},\displaystyle N_{A}(M):=\bigg{\{}\,\sum_{i}u_{i}mv_{i}\ \bigg{|}\ m\in M\bigg{\}},\quad{}_{N_{A}}M:=\left\{\,m\in M\ \bigg{|}\ \sum_{i}u_{i}mv_{i}=0\right\},
IA(M):={j(miν1(ai)aimi)|miM,aiA}.\displaystyle I_{A}(M):=\left\{\,\sum_{j}(m_{i}\nu^{-1}(a_{i})-a_{i}m_{i})\ \bigg{|}\ m_{i}\in M,a_{i}\in A\right\}.

The vector spaces (MNA)/IA(M)({}_{N_{A}}M)/I_{A}(M) and MA/NA(M)M^{A}/N_{A}(M) appear in the following exact sequence

(2.3) 0MNA/IA(M)H0(A,Mν11)N¯H0(A,M)MA/NA(M)0,\displaystyle 0\rightarrow{}_{N_{A}}M/I_{A}(M)\rightarrow\operatorname{H}\nolimits_{0}(A,{}_{1}M_{\nu^{-1}})\xrightarrow{\overline{N}}\operatorname{H}\nolimits^{0}(A,M)\rightarrow M^{A}/N_{A}(M)\rightarrow 0,

where N¯\overline{N} is induced by the norm map N:Mν11MN:{}_{1}M_{\nu^{-1}}\rightarrow M defined in [Nak57, San92] which sends mMν11m\in{}_{1}M_{\nu^{-1}} to N(m)=iuimviN(m)=\sum_{i}u_{i}mv_{i}. In order to prove the existence of the exact sequence (2.3)(\ref{eq:3}), without loss of generality, we may suppose that the beginning of 𝐓\operatorname{\mathbf{T}}\nolimits is of the form

where the maps above are given by as follows:

d1:A3A2;abcabcabc,\displaystyle d_{1}:A^{\otimes 3}\rightarrow A^{\otimes 2};\quad a\otimes b\otimes c\mapsto ab\otimes c-a\otimes bc,
ε:A2A;xyxy,\displaystyle\varepsilon:A^{\otimes 2}\rightarrow A;\quad x\otimes y\mapsto xy,
η:A(A2)ν1;xiuiν(x)vi,\displaystyle\eta:A\rightarrow{}_{1}(A^{\otimes 2})_{\nu};\quad x\mapsto\sum_{i}u_{i}\nu(x)\otimes v_{i},
d1:(A2)ν1((A3))ν1;xy[1a1axyxya].\displaystyle d_{-1}:{}_{1}(A^{\otimes 2})_{\nu}\rightarrow{}_{1}((A^{\otimes 3})^{\vee})_{\nu};\quad x\otimes y\mapsto[1\otimes a\otimes 1\mapsto ax\otimes y-x\otimes ya].

A direct calculation shows that there exists a morphism N¯:H0(A,Mν11)H0(A,M)\overline{N}:\operatorname{H}\nolimits_{0}(A,{}_{1}M_{\nu^{-1}})\rightarrow\operatorname{H}\nolimits^{0}(A,M) making the following square commute:

Mν11\textstyle{{}_{1}M_{\nu^{-1}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}N\scriptstyle{N}M\textstyle{M}H0(A,Mν11)\textstyle{\operatorname{H}\nolimits_{0}(A,{}_{1}M_{\nu^{-1}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}N¯\scriptstyle{\overline{N}}H0(A,M)\textstyle{\operatorname{H}\nolimits^{0}(A,M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

Since N:Mν11MN:{}_{1}M_{\nu^{-1}}\rightarrow M is the 0-th differential of 𝓂Ae(𝐓,M)\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,M), we get the isomorphisms (2.2) and the exact sequence (2.3).

Using Lemma 2.9, one analogously checks that the Tate-Hochschild homology groups H^(A,M)\operatorname{\widehat{H}}\nolimits_{*}(A,M) can be written as follows:

H^n(A,M)={Hn(A,M) if n>0,Hn1(A,Mν1) if n<1,\displaystyle\operatorname{\widehat{H}}\nolimits_{n}(A,M)=\begin{cases}\operatorname{H}\nolimits_{n}(A,M)&\mbox{ if }n>0,\\[3.0pt] \operatorname{H}\nolimits^{-n-1}(A,{}_{1}M_{\nu})&\mbox{ if }n<-1,\end{cases}

and there exists an exact sequence

0H^0(A,M)H0(A,M)N¯H0(A,Mν1)H^1(A,M)0,\displaystyle 0\rightarrow\operatorname{\widehat{H}}\nolimits_{0}(A,M)\rightarrow\operatorname{H}\nolimits_{0}(A,M)\xrightarrow{\overline{N^{\prime}}}\operatorname{H}\nolimits^{0}(A,{}_{1}M_{\nu})\rightarrow\operatorname{\widehat{H}}\nolimits_{-1}(A,M)\rightarrow 0,

where N¯\overline{N^{\prime}} is induced by the morphism N:MMν1N^{\prime}:M\rightarrow{}_{1}M_{\nu} sending mMm\in M to

iuimν(vi)Mν1.\displaystyle\sum_{i}u_{i}m\nu(v_{i})\in{}_{1}M_{\nu}.

Therefore, we have an isomorphisms

H^n(A,M)H^n1(A,Mν11)\displaystyle\operatorname{\widehat{H}}\nolimits^{n}(A,M)\cong\operatorname{\widehat{H}}\nolimits_{-n-1}(A,{}_{1}M_{\nu^{-1}})

for nn\in\operatorname{\mathbb{Z}}\nolimits and an AA-bimodule MM.

We end this section by recalling the definition of weakly projective bimodules in the sense of Sanada [San92]. For bimodules MM and NN over a Frobenius algebra AA, the space Hom,A(M,N)\operatorname{Hom}\nolimits_{-,A}(M,N) of morphisms of right AA-modules becomes an AA-bimodule by defining

(agb)(m):=ag(bm)(agb)(m):=ag(bm)

for a,bA,gHom,A(M,N)a,b\in A,g\in\operatorname{Hom}\nolimits_{-,A}(M,N) and mMm\in M. Similarly, the space HomA,(M,N)\operatorname{Hom}\nolimits_{A,-}(M,N) of morphisms of left AA-modules becomes an AA-bimodule by defining

(agb)(m):=g(ma)b(agb)(m):=g(ma)b

for a,bA,gHomA,(M,N)a,b\in A,g\in\operatorname{Hom}\nolimits_{A,-}(M,N) and mMm\in M. Then we see that

iuigviHomAe(M,N)\sum_{i}u_{i}gv_{i}\in\operatorname{Hom}\nolimits_{A^{\textrm{e}}}(M,N)

for all gHom,A(M,N)g\in\operatorname{Hom}\nolimits_{-,A}(M,N), or gHomA,(M,N)g\in\operatorname{Hom}\nolimits_{A,-}(M,N), where {ui}i\{u_{i}\}_{i} and {vi}i\{v_{i}\}_{i} are dual bases of AA. We say that an AA-bimoudle MM is weakly projective if there exists either gHom,A(M,M)g\in\operatorname{Hom}\nolimits_{-,A}(M,M) or gHomA,(M,M)g\in\operatorname{Hom}\nolimits_{A,-}(M,M) such that

iuigvi=idMEndAe(M).\sum_{i}u_{i}gv_{i}=\operatorname{id}\nolimits_{M}\in\operatorname{End}\nolimits_{A^{\textrm{e}}}(M).

For an AA-bimodule MM, Sanada provided four exact sequences of AA-bimodules which split as exact sequences of one-sided AA-modules and whose middle terms are weakly projective:

0K(M)AMM0(right A-splitting),\displaystyle 0\rightarrow K(M)\rightarrow A\otimes M\rightarrow M\rightarrow 0\quad\mbox{(right $A$-splitting),}
0K(M)MAM0(left A-splitting),\displaystyle 0\rightarrow K^{\prime}(M)\rightarrow M\otimes A\rightarrow M\rightarrow 0\quad\mbox{(left $A$-splitting),}
0MHomk(AA,MA)C(M)0(right A-splitting),\displaystyle 0\rightarrow M\rightarrow\operatorname{Hom}\nolimits_{k}(A_{A},M_{A})\rightarrow C(M)\rightarrow 0\quad\mbox{(right $A$-splitting),}
0MHomk(AA,MA)C(M)0(left A-splitting).\displaystyle 0\rightarrow M\rightarrow\operatorname{Hom}\nolimits_{k}({}_{A}A,{}_{A}M)\rightarrow C^{\prime}(M)\rightarrow 0\quad\mbox{(left $A$-splitting).}
Lemma 2.11 ([San92, Lemma 1.3]).

Let AA be a Frobenius algebra. If an AA-bimodule MM is weakly projective, then H^i(A,M)\operatorname{\widehat{H}}\nolimits^{i}(A,M) vanishes for all ii\in\operatorname{\mathbb{Z}}\nolimits.

Corollary 2.12.

Let AA be a Frobenius algebra. If an AA-bimodule MM is weakly projective, then H^i(A,M)\operatorname{\widehat{H}}\nolimits_{i}(A,M) vanishes for all ii\in\operatorname{\mathbb{Z}}\nolimits.

Proof.

Let MM be a weakly projective AA-bimodule. Observe that Mβα{}_{\alpha}M_{\beta} is also weakly projective for any α,βAut(A)\alpha,\beta\in\operatorname{Aut}\nolimits(A). Therefore, the statement follows from the fact that there exists an isomorphism

H^i(A,M)H^i1(A,Mν1)\operatorname{\widehat{H}}\nolimits_{i}(A,M)\cong\operatorname{\widehat{H}}\nolimits^{-i-1}(A,{}_{1}M_{\nu})

for any ii\in\operatorname{\mathbb{Z}}\nolimits, where ν\nu is the Nakayama automorphism of AA. ∎

Each of the four exact sequences above yields a long exact sequence of Tate-Hochschild (co)homology groups with connecting homomorphisms \partial, so that we have the following.

Corollary 2.13 ([San92, Corollary 1.5]).

for any ii\in\operatorname{\mathbb{Z}}\nolimits, there exist isomorphisms

:H^i(A,M)H^i+1(A,K(M))\displaystyle\partial:\operatorname{\widehat{H}}\nolimits^{i}(A,M)\xrightarrow{\ \sim\ }\operatorname{\widehat{H}}\nolimits^{i+1}(A,K(M))\quad (orH^i+1(A,K(M))),\displaystyle(or\xrightarrow{\ \sim\ }\operatorname{\widehat{H}}\nolimits^{i+1}(A,K^{\prime}(M))),
1:H^i(A,M)H^i1(A,C(M))\displaystyle\partial^{-1}:\operatorname{\widehat{H}}\nolimits^{i}(A,M)\xrightarrow{\ \sim\ }\operatorname{\widehat{H}}\nolimits^{i-1}(A,C(M))\quad (orH^i1(A,C(M))),\displaystyle(or\xrightarrow{\ \sim\ }\operatorname{\widehat{H}}\nolimits^{i-1}(A,C^{\prime}(M))),
:H^i(A,M)H^i1(A,K(M))\displaystyle\partial:\operatorname{\widehat{H}}\nolimits_{i}(A,M)\xrightarrow{\ \sim\ }\operatorname{\widehat{H}}\nolimits_{i-1}(A,K(M))\quad (orH^i1(A,K(M))),\displaystyle(or\xrightarrow{\ \sim\ }\operatorname{\widehat{H}}\nolimits_{i-1}(A,K^{\prime}(M))),
1:H^i(A,M)H^i+1(A,C(M))\displaystyle\partial^{-1}:\operatorname{\widehat{H}}\nolimits_{i}(A,M)\xrightarrow{\ \sim\ }\operatorname{\widehat{H}}\nolimits_{i+1}(A,C(M))\quad (orH^i+1(A,C(M))).\displaystyle(or\xrightarrow{\ \sim\ }\operatorname{\widehat{H}}\nolimits_{i+1}(A,C^{\prime}(M))).

3. Cup product and cap product in Tate-Hochschild theory

Our aim in this section is to prove our main theorem. For this purpose, we define cup product and cap product on Tate-Hochschild (co)homology in an analogous way to the discussion in [Bro94]. As in Hochschild theory, we also give two certain products, called composition products, which are equivalent to the cup product and the cap product, respectively. Throughout this section, let AA denote a finite dimensional Frobenius algebra over a field kk, unless otherwise stated.

3.1. Cup product and cap product

In this subsection, we show the existence of diagonal approximation for arbitrary complete resolution of AA and prove that all diagonal approximations define exactly one cup product and one cap product.

Recall that the cup product on Hochschild cohomology groups is defined by using a diagonal approximation Δ:𝐏𝐏A𝐏\Delta:\operatorname{\mathbf{P}}\nolimits\rightarrow\operatorname{\mathbf{P}}\nolimits\otimes_{A}\operatorname{\mathbf{P}}\nolimits for a single projective resolution 𝐏\operatorname{\mathbf{P}}\nolimits of AA. We will define cup product on Tate-Hochschild cohomology groups in a similar way. However, we fail to develop the theory of cup product when we use the ordinary tensor product, because we need to define the cup product of two elements of degree ii and of degree jj with i,ji,j\in\operatorname{\mathbb{Z}}\nolimits. For this, we must consider the complete tensor product of two chain complexes.

Definition 3.1.

Let CC be a chain complex of right modules over a ((not necessarily Frobenius)) algebra AA and CC^{\prime} a chain complex of left AA-modules. Then the complete tensor product C^ACC\operatorname{\widehat{\otimes}}\nolimits_{A}C^{\prime} is defined to be the chain complex with components

(C^AC)n:=iCniACi(C\operatorname{\widehat{\otimes}}\nolimits_{A}C^{\prime})_{n}:=\prod_{i\in\operatorname{\mathbb{Z}}\nolimits}C_{n-i}\otimes_{A}C^{\prime}_{i}

and differentials

dnC^AC((xniAxi)i):=(dC(xni)Axi+(1)nixniAdC(xi))id_{n}^{C\operatorname{\widehat{\otimes}}\nolimits_{A}C^{\prime}}\left((x_{n-i}\otimes_{A}x^{\prime}_{i})_{i\in\operatorname{\mathbb{Z}}\nolimits}\right):=\left(d^{C}(x_{n-i})\otimes_{A}x^{\prime}_{i}+(-1)^{n-i}x_{n-i}\otimes_{A}d^{C^{\prime}}(x^{\prime}_{i})\right)_{i\in\operatorname{\mathbb{Z}}\nolimits}

for any (xniAxi)i(C^AC)n(x_{n-i}\otimes_{A}x^{\prime}_{i})_{i\in\operatorname{\mathbb{Z}}\nolimits}\in(C\operatorname{\widehat{\otimes}}\nolimits_{A}C^{\prime})_{n}.

Definition 3.2.

Let C1C_{1} and C1C^{\prime}_{1} be chain complexes of right modules over a ((not necessarily Frobenius)) algebra AA and C2C_{2} and C2C^{\prime}_{2} chain complexes of left AA-modules. Let u:C1C1u:C_{1}\rightarrow C^{\prime}_{1} and v:C2C2v:C_{2}\rightarrow C^{\prime}_{2} be graded AA-linear maps of degree ss and of degree tt. The complete tensor product u^Avu\operatorname{\widehat{\otimes}}\nolimits_{A}v of uu with vv over AA is defined as the AA-linear graded map of degree s+ts+t defined by

(u^Av)((xniyi)i):=((1)t(ni)u(xni)Av(yi))i(u\operatorname{\widehat{\otimes}}\nolimits_{A}v)\left((x_{n-i}\otimes y_{i})_{i\in\operatorname{\mathbb{Z}}\nolimits}\right):=\left((-1)^{t(n-i)}u(x_{n-i})\otimes_{A}v(y_{i})\right)_{i\in\operatorname{\mathbb{Z}}\nolimits}

for any (xniyi)i(C1^AC2)n(x_{n-i}\otimes y_{i})_{i\in\operatorname{\mathbb{Z}}\nolimits}\in(C_{1}\operatorname{\widehat{\otimes}}\nolimits_{A}C_{2})_{n}. The complete tensor product of graded maps induces a chain map

𝓂A(C1,C1)𝓂A(C2,C2)𝓂A(C1^AC2,C1^AC2).\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A}(C_{1},C^{\prime}_{1})\otimes\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A}(C_{2},C^{\prime}_{2})\rightarrow\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A}(C_{1}\operatorname{\widehat{\otimes}}\nolimits_{A}C_{2},C^{\prime}_{1}\operatorname{\widehat{\otimes}}\nolimits_{A}C^{\prime}_{2}).

Remark that we can rewrite the differential dC^ACd^{C\operatorname{\widehat{\otimes}}\nolimits_{A}C^{\prime}} defined above as

dC^AC=dC^Aid+id^AdC.d^{C\operatorname{\widehat{\otimes}}\nolimits_{A}C^{\prime}}=d^{C}\operatorname{\widehat{\otimes}}\nolimits_{A}\operatorname{id}\nolimits+\operatorname{id}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A}d^{C^{\prime}}.

Let 𝐓\operatorname{\mathbf{T}}\nolimits be a complete resolution of AA with augmentation ε:𝐓A\varepsilon:\operatorname{\mathbf{T}}\nolimits\rightarrow A. We say that a chain map Δ^:𝐓𝐓^A𝐓\widehat{\Delta}:\operatorname{\mathbf{T}}\nolimits\rightarrow\operatorname{\mathbf{T}}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A}\operatorname{\mathbf{T}}\nolimits is a diagonal approximation if it satisfies (ε^Aε)Δ^=ε(\varepsilon\operatorname{\widehat{\otimes}}\nolimits_{A}\varepsilon)\widehat{\Delta}=\varepsilon. Note that 𝐓^A𝐓\operatorname{\mathbf{T}}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A}\operatorname{\mathbf{T}}\nolimits is no longer a complete resolution of AA, because all the components of 𝐓^A𝐓\operatorname{\mathbf{T}}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A}\operatorname{\mathbf{T}}\nolimits are not finitely generated. Nevertheless, we prove that there exists a diagonal approximation for arbitrary complete resolution of AA. For this purpose, we need the following three lemmas.

Lemma 3.3.

Let 𝐓\operatorname{\mathbf{T}}\nolimits be a complete resolution of AA. Then 𝐓^A𝐓\operatorname{\mathbf{T}}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A}\operatorname{\mathbf{T}}\nolimits is acyclic and dimensionwise projective as AA-bimodules.

Proof.

First, we will show the projectivity of each component of 𝐓^A𝐓\operatorname{\mathbf{T}}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A}\operatorname{\mathbf{T}}\nolimits. This follows from our assumption that AA is Frobenius and the fact that AeAAeA^{\textrm{e}}\otimes_{A}A^{\textrm{e}} is projective over AeA^{\textrm{e}}. In order to prove the acyclicity of 𝐓^A𝐓\operatorname{\mathbf{T}}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A}\operatorname{\mathbf{T}}\nolimits, we construct a contracting homotopy for 𝐓^A𝐓\operatorname{\mathbf{T}}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A}\operatorname{\mathbf{T}}\nolimits as a complex of right AA-modules. Since any complete resolution 𝐓\operatorname{\mathbf{T}}\nolimits of AA is contractible as a complex of left AA-modules, we obtain

a contracting homotopy hh. Set H=id𝐓^Ah:𝐓^A𝐓𝐓^A𝐓H=\operatorname{id}\nolimits_{\operatorname{\mathbf{T}}\nolimits}\operatorname{\widehat{\otimes}}\nolimits_{A}h:\operatorname{\mathbf{T}}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A}\operatorname{\mathbf{T}}\nolimits\rightarrow\operatorname{\mathbf{T}}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A}\operatorname{\mathbf{T}}\nolimits. Then HH is a contracting homotopy for 𝐓^A𝐓\operatorname{\mathbf{T}}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A}\operatorname{\mathbf{T}}\nolimits. Indeed, we have

d𝐓^A𝐓H\displaystyle d^{\operatorname{\mathbf{T}}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A}\operatorname{\mathbf{T}}\nolimits}H +Hd𝐓^A𝐓\displaystyle+Hd^{\operatorname{\mathbf{T}}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A}\operatorname{\mathbf{T}}\nolimits}
=(d𝐓^Aid+id^d𝐓)(id^h)+(id^h)(d𝐓^Aid+id^d𝐓)\displaystyle=(d^{\operatorname{\mathbf{T}}\nolimits}\operatorname{\widehat{\otimes}}\nolimits_{A}\operatorname{id}\nolimits+\operatorname{id}\nolimits\operatorname{\widehat{\otimes}}\nolimits d^{\operatorname{\mathbf{T}}\nolimits})(\operatorname{id}\nolimits\operatorname{\widehat{\otimes}}\nolimits h)+(\operatorname{id}\nolimits\operatorname{\widehat{\otimes}}\nolimits h)(d^{\operatorname{\mathbf{T}}\nolimits}\operatorname{\widehat{\otimes}}\nolimits_{A}\operatorname{id}\nolimits+\operatorname{id}\nolimits\operatorname{\widehat{\otimes}}\nolimits d^{\operatorname{\mathbf{T}}\nolimits})
=d𝐓^Ah+id^Ad𝐓hd𝐓^Ah+id^Ahd𝐓\displaystyle=d^{\operatorname{\mathbf{T}}\nolimits}\operatorname{\widehat{\otimes}}\nolimits_{A}h+\operatorname{id}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A}d^{\operatorname{\mathbf{T}}\nolimits}h-d^{\operatorname{\mathbf{T}}\nolimits}\operatorname{\widehat{\otimes}}\nolimits_{A}h+\operatorname{id}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A}hd^{\operatorname{\mathbf{T}}\nolimits}
=id^A(d𝐓h+hd𝐓)=id^Aid.\displaystyle=\operatorname{id}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A}(d^{\operatorname{\mathbf{T}}\nolimits}h+hd^{\operatorname{\mathbf{T}}\nolimits})=\operatorname{id}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A}\operatorname{id}\nolimits.

This completes the proof. ∎

Lemma 3.4.

Let CC and CC^{\prime} be acyclic chain complexes over a ((not necessarily Frobenius)) algebra AA. Assume that CiC_{i} is projective over AA for i1i\geq 1 and that CiC^{\prime}_{i} is injective over AA for i1i\leq-1. If there exists a morphism τ0:C0C0\tau_{0}:C_{0}\rightarrow C^{\prime}_{0} satisfies d0Cτ0d1C=0d^{C^{\prime}}_{0}\tau_{0}d^{C}_{1}=0, then τ0\tau_{0} extends to a chain map τ:CC\tau:C\rightarrow C^{\prime}, up to homotopy.

Proof.

Consider a commutative diagram

\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}C2\textstyle{C_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d2C\scriptstyle{d^{C}_{2}}C1\textstyle{C_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d1C\scriptstyle{d^{C}_{1}}C0\textstyle{C_{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d0C\scriptstyle{d^{C}_{0}}τ0\scriptstyle{\tau_{0}}Cokerd1C\textstyle{\operatorname{Coker}\nolimits d^{C}_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d0Cτ0¯\scriptstyle{\overline{d^{C^{\prime}}_{0}\tau_{0}}}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\cdots}\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}C2\textstyle{C^{\prime}_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d2C\scriptstyle{d^{C^{\prime}}_{2}}C1\textstyle{C^{\prime}_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d1C\scriptstyle{d^{C^{\prime}}_{1}}C0\textstyle{C^{\prime}_{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d0C\scriptstyle{d^{C^{\prime}}_{0}}C1\textstyle{C^{\prime}_{-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\cdots}

where the morphism d0Cτ0¯\overline{d^{C^{\prime}}_{0}\tau_{0}} is given by x0¯d0Cτ0(x0)\overline{x_{0}}\mapsto d^{C^{\prime}}_{0}\tau_{0}(x_{0}) for x0¯Cokerd1C\overline{x_{0}}\in\operatorname{Coker}\nolimits d^{C}_{1}. It follows form Lemma 1.1 that there uniquely (up to homotopy) exists a family {τi:CiCi}i0\{\tau_{i}:C_{i}\rightarrow C^{\prime}_{i}\}_{i\geq 0} such that diCτi=τi1diCd^{C^{\prime}}_{i}\tau_{i}=\tau_{i-1}d^{C}_{i} for i1i\geq 1. Applying Lemma 1.2 to a commutative diagram

\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}C2\textstyle{C_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d2C\scriptstyle{d^{C}_{2}}τ2\scriptstyle{\tau_{2}}C1\textstyle{C_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d1C\scriptstyle{d^{C}_{1}}τ1\scriptstyle{\tau_{1}}C0\textstyle{C_{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d0C\scriptstyle{d^{C}_{0}}τ0\scriptstyle{\tau_{0}}C1\textstyle{C_{-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d1C\scriptstyle{d^{C}_{-1}}C2\textstyle{C_{-2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\cdots}\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}C2\textstyle{C^{\prime}_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d2C\scriptstyle{d^{C^{\prime}}_{2}}C1\textstyle{C^{\prime}_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d1C\scriptstyle{d^{C^{\prime}}_{1}}C0\textstyle{C^{\prime}_{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d0C\scriptstyle{d^{C^{\prime}}_{0}}C1\textstyle{C^{\prime}_{-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d1C\scriptstyle{d^{C^{\prime}}_{-1}}C2\textstyle{C^{\prime}_{-2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\cdots}

we have that the family {τi}i0\{\tau_{i}\}_{i\geq 0} extends to a chain map τ:CC\tau:C\rightarrow C^{\prime}, which is uniquely determined up to homotopy. ∎

Lemma 3.5.

Let PP be a finitely generated projective bimoudle over a ((not necessarily Frobenius)) algebra AA and CC an acyclic chain complex of AA-bimodules. Assume that CC is contractible as a complex of left ((resp., right)) AA-modules. Then CAP(C\otimes_{A}P\ (resp., PAC)P\otimes_{A}C) is contractible as a complex of AA-bimodules.

Proof.

We prove the statement only for the case that CC is contractible as a complex of left AA-modules. It suffices to show the statement for P=AeP=A^{\textrm{e}}. We construct a contracting homotopy for CAAeC\otimes_{A}A^{\textrm{e}}. Let hh be a contracting homotopy for CA{}_{A}C. A direct computation shows that the graded map hidA:CACAh\otimes\operatorname{id}\nolimits_{A}:C\otimes A\rightarrow C\otimes A of AA-bimodules of degree 11 is a contracting homotopy. Since there is an isomorphism CAAeCAC\otimes_{A}A^{\textrm{e}}\cong C\otimes A of chain complexes of AA-bimodules, the graded map induced by hidAh\otimes\operatorname{id}\nolimits_{A} via this isomorphism is a contracting homotopy for CAAeC\otimes_{A}A^{\textrm{e}}. ∎

We are now able to show the existence of a diagonal approximation for any complete resolution of a Frobenius algebra.

Theorem 3.6.

Let 𝐓\operatorname{\mathbf{T}}\nolimits be a complete resolution of AA. Then there uniquely ((up to homotopy)) exists a diagonal approximation Δ^:𝐓𝐓^A𝐓\widehat{\Delta}:\operatorname{\mathbf{T}}\nolimits\rightarrow\operatorname{\mathbf{T}}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A}\operatorname{\mathbf{T}}\nolimits.

Proof.

In view of Lemmas 3.3 and 3.4, it suffices to construct a map τ:T0i(TiATi)\tau:T_{0}\rightarrow\prod_{i\in\operatorname{\mathbb{Z}}\nolimits}(T_{i}\otimes_{A}T_{-i}) such that (1)d0𝐓^A𝐓τd1𝐓=0(1)\,d^{\operatorname{\mathbf{T}}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A}\operatorname{\mathbf{T}}\nolimits}_{0}\,\tau\,d^{\operatorname{\mathbf{T}}\nolimits}_{1}=0 and (2)(ε^Aε)τ=ε(2)\,(\varepsilon\operatorname{\widehat{\otimes}}\nolimits_{A}\varepsilon)\tau=\varepsilon, where ε:𝐓A\varepsilon:\operatorname{\mathbf{T}}\nolimits\rightarrow A is the augmentation of 𝐓\operatorname{\mathbf{T}}\nolimits. Let τr:T0TrATr\tau_{r}:T_{0}\rightarrow T_{r}\otimes_{A}T_{-r} denote the composition of τ\tau with the rr-th canonical projection on i(TiATi)\prod_{i\in\operatorname{\mathbb{Z}}\nolimits}(T_{i}\otimes_{A}T_{-i}). Set

di,j:=di𝐓AidTjanddi,j′′:=(1)iidTjAdi𝐓.d^{\prime}_{i,j}:=d^{\operatorname{\mathbf{T}}\nolimits}_{i}\otimes_{A}\operatorname{id}\nolimits_{T_{j}}\quad\mbox{and}\quad d^{\prime\prime}_{i,j}:=(-1)^{i}\operatorname{id}\nolimits_{T_{j}}\otimes_{A}d^{\operatorname{\mathbf{T}}\nolimits}_{i}.

Then we can rewrite the first condition (1)(1) as

(3.1) (di,iτi+di1,i+1′′τi1)|B0(𝐓)=0 for each i.\displaystyle\left(d^{\prime}_{i,-i}\tau_{i}+d^{\prime\prime}_{i-1,-i+1}\tau_{i-1}\right)\big{|}_{B_{0}(\operatorname{\mathbf{T}}\nolimits)}=0\quad\mbox{ for each $i\in\operatorname{\mathbb{Z}}\nolimits$.}

Since T0T_{0} is projective, there exists a morphism τ0:T0T0AT0\tau_{0}:T_{0}\rightarrow T_{0}\otimes_{A}T_{0} of AA-bimodules such that (ε^Aε)τ0=ε(\varepsilon\operatorname{\widehat{\otimes}}\nolimits_{A}\varepsilon)\tau_{0}=\varepsilon. Suppose that i>0i>0 and that we have defined τj\tau_{j} with 0j<i0\leq j<i satisfying the condition (3.1). Consider the following diagram with exact row:

B0(𝐓)\textstyle{B_{0}(\operatorname{\mathbf{T}}\nolimits)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}di1,i+1′′τi1|B0(𝐓)\scriptstyle{-d^{\prime\prime}_{i-1,-i+1}\tau_{i-1}\big{|}_{B_{0}(\operatorname{\mathbf{T}}\nolimits)}}TiATi\textstyle{T_{i}\otimes_{A}T_{-i}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}di,i\scriptstyle{d^{\prime}_{i,-i}}Ti1ATi\textstyle{T_{i-1}\otimes_{A}T_{-i}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}di1,i\scriptstyle{d^{\prime}_{i-1,-i}}Ti2ATi\textstyle{T_{i-2}\otimes_{A}T_{-i}}

It follows from Lemma 3.5 that the complex (𝐓ATi,d,i)(\operatorname{\mathbf{T}}\nolimits\otimes_{A}T_{-i},d^{\prime}_{\bullet,-i}) of AA-bimodules is contractible. Let h=(hi)ih=(h_{i})_{i\in\operatorname{\mathbb{Z}}\nolimits} be a contracting homotopy for 𝐓ATi\operatorname{\mathbf{T}}\nolimits\otimes_{A}T_{-i}. Put

τi:=hi1(di1,i+1′′τi1):T0TiATi.\tau_{i}:=-h_{i-1}(d^{\prime\prime}_{i-1,-i+1}\tau_{i-1}):T_{0}\rightarrow T_{i}\otimes_{A}T_{-i}.

We now claim that di1,i(di1,1i′′τi1)|B0(𝐓)=0d^{\prime}_{i-1,-i}(-d^{\prime\prime}_{i-1,1-i}\tau_{i-1})\big{|}_{B_{0}(\operatorname{\mathbf{T}}\nolimits)}=0 hols for i1i\geq 1. If i=1i=1, then we have on B0(𝐓)B_{0}(\operatorname{\mathbf{T}}\nolimits)

d0,1(d0,0′′τ0)\displaystyle d^{\prime}_{0,-1}(-d^{\prime\prime}_{0,0}\tau_{0}) =(d0Ad0)τ0\displaystyle=-(d_{0}\otimes_{A}d_{0})\tau_{0}
=(ηAη)(εAε)τ0\displaystyle=-(\eta\otimes_{A}\eta)(\varepsilon\otimes_{A}\varepsilon)\tau_{0}
=0.\displaystyle=0.

Assume that i>1i>1. We know that

(di1,1iτi1+di2,2i′′τi2)|B0(𝐓)=0(d^{\prime}_{i-1,1-i}\tau_{i-1}+d^{\prime\prime}_{i-2,2-i}\tau_{i-2})\big{|}_{B_{0}(\operatorname{\mathbf{T}}\nolimits)}=0

holds, so that we have on B0(𝐓)B_{0}(\operatorname{\mathbf{T}}\nolimits)

di1,i(di1,1i′′τi1)\displaystyle d^{\prime}_{i-1,-i}(-d^{\prime\prime}_{i-1,1-i}\tau_{i-1}) =(di1,idi1,1i′′)τi1\displaystyle=(-d^{\prime}_{i-1,-i}d^{\prime\prime}_{i-1,1-i})\tau_{i-1}
=(di2,1i′′di1,1i)τi1\displaystyle=(d^{\prime\prime}_{i-2,1-i}d^{\prime}_{i-1,1-i})\tau_{i-1}
=di2,1i′′(di2,2i′′τi2)\displaystyle=-d^{\prime\prime}_{i-2,1-i}(d^{\prime\prime}_{i-2,2-i}\tau_{i-2})
=0\displaystyle=0

and hence on B0(𝐓)B_{0}(\operatorname{\mathbf{T}}\nolimits)

di,iτi\displaystyle d^{\prime}_{i,-i}\tau_{i} =di,ihi1(di1,i+1′′τi1)\displaystyle=d^{\prime}_{i,-i}h_{i-1}(-d^{\prime\prime}_{i-1,-i+1}\tau_{i-1})
=(idhi2di1,i)(di1,i+1′′τi1)\displaystyle=(\operatorname{id}\nolimits-h_{i-2}d^{\prime}_{i-1,-i})(-d^{\prime\prime}_{i-1,-i+1}\tau_{i-1})
=(di1,i+1′′τi1)+hi2(di1,i(di1,i+1′′τi1))\displaystyle=(-d^{\prime\prime}_{i-1,-i+1}\tau_{i-1})+h_{i-2}(d^{\prime}_{i-1,-i}(d^{\prime\prime}_{i-1,-i+1}\tau_{i-1}))
=di1,i+1′′τi1.\displaystyle=-d^{\prime\prime}_{i-1,-i+1}\tau_{i-1}.

We have constructed {τi}i0\{\tau_{i}\}_{i\geq 0} satisfying the condition (3.1). A dual argument using descending induction on i<0i<0 shows that we get the other components τi\tau_{i} with i<0i<0. In conclusion, the statement follows from Lemmas 3.3 and 3.4. ∎

Remark 3.7.

Let 𝐓\operatorname{\mathbf{T}}\nolimits be a complete resolution of AA. We denote by 𝐏\operatorname{\mathbf{P}}\nolimits the non-negative truncation 𝐓0\operatorname{\mathbf{T}}\nolimits_{\geq 0}. Recall that 𝐏\operatorname{\mathbf{P}}\nolimits is a projective resolution of AA. Set

𝐓^A𝐓0:=n0(0pnTnpATp)\displaystyle\operatorname{\mathbf{T}}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A}\operatorname{\mathbf{T}}\nolimits_{\geq 0}:=\bigoplus_{n\geq 0}\left(\bigoplus_{0\leq p\leq n}T_{n-p}\otimes_{A}T_{p}\right)

and

d𝐓^A𝐓0:=d0𝐓Aid+idAd0𝐓.\displaystyle d^{\operatorname{\mathbf{T}}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A}\operatorname{\mathbf{T}}\nolimits_{\geq 0}}:=d_{\geq 0}^{\operatorname{\mathbf{T}}\nolimits}\otimes_{A}\operatorname{id}\nolimits+\operatorname{id}\nolimits\otimes_{A}d_{\geq 0}^{\operatorname{\mathbf{T}}\nolimits}.

Then (𝐓^A𝐓0,d𝐓^A𝐓0)(\operatorname{\mathbf{T}}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A}\operatorname{\mathbf{T}}\nolimits_{\geq 0},d^{\operatorname{\mathbf{T}}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A}\operatorname{\mathbf{T}}\nolimits_{\geq 0}}) is nothing but a projective resolution 𝐏A𝐏\operatorname{\mathbf{P}}\nolimits\otimes_{A}\operatorname{\mathbf{P}}\nolimits of AA. If Δ^:𝐓𝐓^A𝐓\widehat{\Delta}:\operatorname{\mathbf{T}}\nolimits\rightarrow\operatorname{\mathbf{T}}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A}\operatorname{\mathbf{T}}\nolimits is a diagonal approximation, then it can be decomposed as

Δ^=n(pΔ^p(n))\displaystyle\widehat{\Delta}=\prod_{n\in\operatorname{\mathbb{Z}}\nolimits}\left(\,\prod_{p\in\operatorname{\mathbb{Z}}\nolimits}\widehat{\Delta}^{(n)}_{p}\right)

with Δ^p(n):TnTnpATp\widehat{\Delta}^{(n)}_{p}:T_{n}\rightarrow T_{n-p}\otimes_{A}T_{p}. We denote by Δ^0\widehat{\Delta}_{\geq 0} a graded AeA^{\textrm{e}}-linear map

n0(0pnΔ^p(n)):𝐓0𝐓^A𝐓0.\displaystyle\prod_{n\geq 0}\left(\prod_{0\leq p\leq n}\widehat{\Delta}^{(n)}_{p}\right):\operatorname{\mathbf{T}}\nolimits_{\geq 0}\rightarrow\operatorname{\mathbf{T}}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A}\operatorname{\mathbf{T}}\nolimits_{\geq 0}.

By the definition of Δ^0\widehat{\Delta}_{\geq 0}, the graded map Δ^0\widehat{\Delta}_{\geq 0} becomes an augmentation-preserving chain map from 𝐏\operatorname{\mathbf{P}}\nolimits to 𝐏A𝐏\operatorname{\mathbf{P}}\nolimits\otimes_{A}\operatorname{\mathbf{P}}\nolimits, which means that Δ^0\widehat{\Delta}_{\geq 0} is a diagonal approximation for the projective resolution 𝐏\operatorname{\mathbf{P}}\nolimits.

Theorem 3.6 allows us to define cup product and cap product on Tate-Hochschild (co)homology groups: for AA-bimodules MM and NN, we define a graded kk-linear map

(3.2) :𝓂Ae(𝐓,M)𝓂Ae(𝐓,N)𝓂Ae(𝐓,MAN)\displaystyle\smile:\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,M)\otimes\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,N)\rightarrow\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,M\otimes_{A}N)

by uv:=(u^Av)Δ^u\smile v:=(u\operatorname{\widehat{\otimes}}\nolimits_{A}v)\widehat{\Delta} for homogeneous elements u𝓂Ae(𝐓,M)u\in\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,M) and v𝓂Ae(𝐓,N)v\in\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,N). One can easily check that \smile is a chain map, so that it induces an operator

(3.3) :H^(A,M)H^(A,N)H^(A,MAN).\displaystyle\smile:\operatorname{\widehat{H}}\nolimits^{*}(A,M)\otimes\operatorname{\widehat{H}}\nolimits^{*}(A,N)\rightarrow\operatorname{\widehat{H}}\nolimits^{*}(A,M\otimes_{A}N).

For uH^r(A,M)u\in\operatorname{\widehat{H}}\nolimits^{r}(A,M) and vH^s(A,N)v\in\operatorname{\widehat{H}}\nolimits^{s}(A,N), we call uvH^r+s(A,MAN)u\smile v\in\operatorname{\widehat{H}}\nolimits^{r+s}(A,M\otimes_{A}N) the cup product of uu and vv. On the other hand, consider the composition

:𝓂Ae(𝐓,M)(𝐓AeN)𝐓Ae(MAN)\displaystyle\frown:\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,M)\otimes(\operatorname{\mathbf{T}}\nolimits\otimes_{A^{\textrm{e}}}N)\rightarrow\operatorname{\mathbf{T}}\nolimits\otimes_{A^{\textrm{e}}}(M\otimes_{A}N)

of two chain maps

id\displaystyle\operatorname{id}\nolimits (Δ^Aeid):\displaystyle\otimes(\widehat{\Delta}\otimes_{A^{\textrm{e}}}\operatorname{id}\nolimits):
𝓂Ae(𝐓,M)(𝐓AeN)𝓂Ae(𝐓,M)((𝐓^A𝐓)AeN)\displaystyle\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,M)\otimes(\operatorname{\mathbf{T}}\nolimits\otimes_{A^{\textrm{e}}}N)\rightarrow\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,M)\otimes((\operatorname{\mathbf{T}}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A}\operatorname{\mathbf{T}}\nolimits)\otimes_{A^{\textrm{e}}}N)

and

γ:𝓂Ae(𝐓,M)((𝐓^A𝐓)AeN)𝐓Ae(MAN)\displaystyle\gamma:\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,M)\otimes((\operatorname{\mathbf{T}}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A}\operatorname{\mathbf{T}}\nolimits)\otimes_{A^{\textrm{e}}}N)\rightarrow\operatorname{\mathbf{T}}\nolimits\otimes_{A^{\textrm{e}}}(M\otimes_{A}N)

given by

u((x^Ay)Aen)(1)|u||x|xAe(u(y)An).\displaystyle u\otimes((x\operatorname{\widehat{\otimes}}\nolimits_{A}y)\otimes_{A^{\textrm{e}}}n)\mapsto(-1)^{|u||x|}x\otimes_{A^{\textrm{e}}}(u(y)\otimes_{A}n).

Then the composition \frown induces an operator

(3.4) :H^r(A,M)H^s(A,N)H^sr(A,MAN).\displaystyle\frown:\operatorname{\widehat{H}}\nolimits^{r}(A,M)\otimes\operatorname{\widehat{H}}\nolimits_{s}(A,N)\rightarrow\operatorname{\widehat{H}}\nolimits_{s-r}(A,M\otimes_{A}N).

For uH^r(A,M)u\in\operatorname{\widehat{H}}\nolimits^{r}(A,M) and zH^s(A,N)z\in\operatorname{\widehat{H}}\nolimits_{s}(A,N), we call uzH^sr(A,MAN)u\frown z\in\operatorname{\widehat{H}}\nolimits_{s-r}(A,M\otimes_{A}N) the cap product of uu and zz.

Now, we will show the uniqueness of the cup product \smile and of the cap product \frown, that is, each of them does not depend on the choice of a complete resolution and a diagonal approximation. First, we will deal with the cup product.

Proposition 3.8.

The cup product \smile satisfies the following three properties.

  1. (PI)

    Let MM and NN be AA-bimodules. Then there exists a commutative square

    H^0(A,M)H^0(A,N)\textstyle{\operatorname{\widehat{H}}\nolimits^{0}(A,M)\otimes\operatorname{\widehat{H}}\nolimits^{0}(A,N)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\smile}H^0(A,MAN)\textstyle{\operatorname{\widehat{H}}\nolimits^{0}(A,M\otimes_{A}N)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}MA/NA(M)NA/NA(N)\textstyle{M^{A}/N_{A}(M)\otimes N^{A}/N_{A}(N)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(MAN)A/NA(MAN)\textstyle{(M\otimes_{A}N)^{A}/N_{A}(M\otimes_{A}N)}

    where the vertical morphisms are isomorphisms in (2.2)(\ref{eq:5}) and the morphism in the bottom row is given by

    (m+NA(M))(n+NA(N))mAn+NA(MAN).(m+N_{A}(M))\otimes(n+N_{A}(N))\mapsto m\otimes_{A}n+N_{A}(M\otimes_{A}N).
  2. (PII1)

    Let

    0M1M2M30,\displaystyle 0\rightarrow M_{1}\rightarrow M_{2}\rightarrow M_{3}\rightarrow 0,
    0M1ANM2ANM3AN0\displaystyle 0\rightarrow M_{1}\otimes_{A}N\rightarrow M_{2}\otimes_{A}N\rightarrow M_{3}\otimes_{A}N\rightarrow 0

    be exact sequences of AA-bimodules. Then we have

    (γξ)=(γ)ξ\partial(\gamma\smile\xi)=(\partial\gamma)\smile\xi

    for all γH^r(A,M3)\gamma\in\operatorname{\widehat{H}}\nolimits^{r}(A,M_{3}) and ξH^s(A,N)\xi\in\operatorname{\widehat{H}}\nolimits^{s}(A,N), where \partial denotes the connecting homomorphism.

  3. (PII2)

    Let

    0N1N2N30,\displaystyle 0\rightarrow N_{1}\rightarrow N_{2}\rightarrow N_{3}\rightarrow 0,
    0MAN1MAN2MAN30\displaystyle 0\rightarrow M\otimes_{A}N_{1}\rightarrow M\otimes_{A}N_{2}\rightarrow M\otimes_{A}N_{3}\rightarrow 0

    be exact sequences of AA-bimodules. Then we have

    (γξ)=(1)rγ(ξ)\partial(\gamma\smile\xi)=(-1)^{r}\gamma\smile(\partial\xi)

    for all γH^r(A,M)\gamma\in\operatorname{\widehat{H}}\nolimits^{r}(A,M) and ξH^s(A,N3)\xi\in\operatorname{\widehat{H}}\nolimits^{s}(A,N_{3}), where \partial denotes the connecting homomorphism.

Proof.

We only prove that the cup product satisfies (PI) and (PII1); the proof of (PII2) is similar to (PII1). Let 𝐓\operatorname{\mathbf{T}}\nolimits be a complete resolution of AA and Δ^:𝐓𝐓^A𝐓\widehat{\Delta}:\operatorname{\mathbf{T}}\nolimits\rightarrow\operatorname{\mathbf{T}}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A}\operatorname{\mathbf{T}}\nolimits a diagonal approximation. The property that (ε^Aε)Δ^=ε(\varepsilon\operatorname{\widehat{\otimes}}\nolimits_{A}\varepsilon)\widehat{\Delta}=\varepsilon implies that the cup product satisfies (PI).

For (PII1), let α𝓂Ae(𝐓,M)s\alpha\in\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,M)^{s} be a cocycle representing ξH^s(A,N)\xi\in\operatorname{\widehat{H}}\nolimits^{s}(A,N). Then the graded map

α:𝓂Ae(𝐓,M1)𝓂Ae(𝐓,M1AN)-\smile\alpha:\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,M_{1})\rightarrow\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,M_{1}\otimes_{A}N)

induced by α\alpha is a chain map. Thus, we have a commutative diagram of chain complexes with exact rows
0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝓂Ae(𝐓,M1)\textstyle{\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,M_{1})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}α\scriptstyle{-\smile\alpha}𝓂Ae(𝐓,M2)\textstyle{\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,M_{2})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}α\scriptstyle{-\smile\alpha}𝓂Ae(𝐓,M3)\textstyle{\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,M_{3})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}α\scriptstyle{-\smile\alpha}0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝓂Ae(𝐓,M1AN)\textstyle{\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,M_{1}\otimes_{A}N)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝓂Ae(𝐓,M2AN)\textstyle{\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,M_{2}\otimes_{A}N)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝓂Ae(𝐓,M3AN)\textstyle{\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,M_{3}\otimes_{A}N)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}
The property (PII1) follows from the naturality of connecting homomorphism. ∎

Theorem 3.9 ([San92, Theorem 2.1]).

Any two cup products satisfying the three properties (PI)({\rm PI})– (PII2)({\rm PII}_{2}) coincide up to isomorphism.

Let us remark that a system of the properties (PI)({\rm PI})– (PII2)({\rm PII}_{2}) may be originally seen in [Kawse].

As a consequence of the two statements above, we have the following.

Corollary 3.10.

The cup product

:H^(A,M)H^(A,N)H^(A,MAN).\displaystyle\smile:\operatorname{\widehat{H}}\nolimits^{*}(A,M)\otimes\operatorname{\widehat{H}}\nolimits^{*}(A,N)\rightarrow\operatorname{\widehat{H}}\nolimits^{*}(A,M\otimes_{A}N).

does not depend on the choice of a complete resolution and a diagonal approximation.

We know that the cup product satisfies the following properties.

Theorem 3.11 ([San92, Propositions 2.3 and 2.4]).

There exists one diagonal approximation associated with a certain complete resolution 𝐗𝜀A\mathbf{X}\xrightarrow{\varepsilon}A such that

  1. (i)

    The induced cup product \smile is associative, i.e.,

    (u1u2)u3=u1(u2u3)(u_{1}\smile u_{2})\smile u_{3}=u_{1}\smile(u_{2}\smile u_{3})

    for uiH^(A,Mi)u_{i}\in\operatorname{\widehat{H}}\nolimits^{*}(A,M_{i}) with an AA-bimodule MiM_{i}.

  2. (ii)

    The induced cup product \smile endows a graded vector space

    H^(A,A):=iH^i(A,A)=iHi(𝓂Ae(𝐗,A))\operatorname{\widehat{H}}\nolimits^{\bullet}(A,A):=\bigoplus_{i\in\operatorname{\mathbb{Z}}\nolimits}\operatorname{\widehat{H}}\nolimits^{i}(A,A)=\bigoplus_{i\in\operatorname{\mathbb{Z}}\nolimits}\operatorname{H}\nolimits^{i}(\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\mathbf{X},A))

    with a graded commutative algebra structure whose unit is the element represented by the augmentation ε\varepsilon.

Secondly, we deal with the cap product in an analogous way to the cup product.

Proposition 3.12.

The cap product \frown satisfies the following three properties.

  1. (QI)

    Let MM and NN be AA-bimodules. Then there exists a commutative square

    H^0(A,M)H^0(A,N)\textstyle{\operatorname{\widehat{H}}\nolimits^{0}(A,M)\otimes\operatorname{\widehat{H}}\nolimits_{0}(A,N)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\frown}H^0(A,MAN)\textstyle{\operatorname{\widehat{H}}\nolimits_{0}(A,M\otimes_{A}N)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}MA/NA(M)NNA/IA(N)\textstyle{M^{A}/N_{A}(M)\otimes{}_{N_{A}}N/I_{A}(N)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(MAN)NA/IA(MAN)\textstyle{{}_{N_{A}}(M\otimes_{A}N)/I_{A}(M\otimes_{A}N)}

    where the vertical morphisms are isomorphisms in (2.2)(\ref{eq:5}) and the morphism in the bottom row is given by

    (m+NA(M))(n+IA(N))mAn+IA(MAN).(m+N_{A}(M))\otimes(n+I_{A}(N))\mapsto m\otimes_{A}n+I_{A}(M\otimes_{A}N).
  2. (QII1)

    Let

    0M1M2M30,\displaystyle 0\rightarrow M_{1}\rightarrow M_{2}\rightarrow M_{3}\rightarrow 0,
    0M1ANM2ANM3AN0\displaystyle 0\rightarrow M_{1}\otimes_{A}N\rightarrow M_{2}\otimes_{A}N\rightarrow M_{3}\otimes_{A}N\rightarrow 0

    be exact sequences of AA-bimodules. Then we have

    (γξ)=(γ)ξ\partial(\gamma\frown\xi)=(\partial\gamma)\frown\xi

    for all γH^r(A,M3)\gamma\in\operatorname{\widehat{H}}\nolimits^{r}(A,M_{3}) and ξH^s(A,N)\xi\in\operatorname{\widehat{H}}\nolimits_{s}(A,N), where \partial denotes the connecting homomorphism.

  3. (QII2)

    Let

    0N1N2N30,\displaystyle 0\rightarrow N_{1}\rightarrow N_{2}\rightarrow N_{3}\rightarrow 0,
    0MAN1MAN2MAN30\displaystyle 0\rightarrow M\otimes_{A}N_{1}\rightarrow M\otimes_{A}N_{2}\rightarrow M\otimes_{A}N_{3}\rightarrow 0

    be exact sequences of AA-bimodules. Then we have

    (γξ)=(1)rγ(ξ)\partial(\gamma\frown\xi)=(-1)^{r}\gamma\frown(\partial\xi)

    for all γH^r(A,M)\gamma\in\operatorname{\widehat{H}}\nolimits^{r}(A,M) and ξH^s(A,N3)\xi\in\operatorname{\widehat{H}}\nolimits_{s}(A,N_{3}), where \partial denotes the connecting homomorphism.

Proof.

The proof of this proposition is similar to that of Proposition 3.8. ∎

Theorem 3.13.

There exists only one cap product satisfying the three properties (QI)({\rm QI})– (QII2)({\rm QII}_{2}).

Proof.

Let \frown and \frown^{\prime} be any two cap products, and let r,sr,s be arbitrary integers. The property (QI)(\textrm{QI}) implies that \frown coincides with \frown^{\prime} in the case r=s=0r=s=0. A dimension-shifting argument analogue to that in [San92, pp. 78–79] yields that \frown agrees with \frown^{\prime} for any r,sr,s\in\operatorname{\mathbb{Z}}\nolimits. ∎

3.2. Composition products

In this subsection, we will show, as in Hochschild theory, that the cup product and the cap product on Tate-Hochschild (co)homology coincide with some composition products. The following proposition is crucial.

Proposition 3.14.

Let 𝐓\operatorname{\mathbf{T}}\nolimits be an acyclic chain complex of ((not necessarily finitely generated)) projective AA-bimodules, 𝐓εA\operatorname{\mathbf{T}}\nolimits^{\prime}\xrightarrow{\varepsilon^{\prime}}A a complete resolution and MM an AA-bimodule. Then the following statements hold.

  1. (1)

    A chain map εAidM:𝐓AMM\varepsilon^{\prime}\otimes_{A}\operatorname{id}\nolimits_{M}:\operatorname{\mathbf{T}}\nolimits^{\prime}\otimes_{A}M\rightarrow M induces a quasi-isomorphism

    𝓂Ae(𝐓,εAidM):𝓂Ae(𝐓,𝐓AM)𝓂Ae(𝐓,M).\displaystyle\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,\varepsilon^{\prime}\otimes_{A}\operatorname{id}\nolimits_{M}):\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,\operatorname{\mathbf{T}}\nolimits^{\prime}\otimes_{A}M)\rightarrow\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,M).
  2. (2)

    If each component of 𝐓\operatorname{\mathbf{T}}\nolimits is finitely generated, then εAidM\varepsilon^{\prime}\otimes_{A}\operatorname{id}\nolimits_{M} induces a quasi-isomorphism

    id𝐓^Ae(εAidM):𝐓^Ae(𝐓AM)𝐓AeM.\operatorname{id}\nolimits_{\operatorname{\mathbf{T}}\nolimits}\operatorname{\widehat{\otimes}}\nolimits_{A^{\textrm{e}}}(\varepsilon^{\prime}\otimes_{A}\operatorname{id}\nolimits_{M}):\operatorname{\mathbf{T}}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits^{\prime}\otimes_{A}M)\rightarrow\operatorname{\mathbf{T}}\nolimits\otimes_{A^{\textrm{e}}}M.

In particular, if 𝐓\operatorname{\mathbf{T}}\nolimits is a compete resolution of AA, then we have isomorphisms

H^(A,M)H(𝓂Ae(𝐓,𝐓AM)),\displaystyle\operatorname{\widehat{H}}\nolimits^{*}(A,M)\cong\operatorname{H}\nolimits^{*}(\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,\operatorname{\mathbf{T}}\nolimits^{\prime}\otimes_{A}M)),
H^(A,M)H(𝐓^Ae(𝐓AM).\displaystyle\operatorname{\widehat{H}}\nolimits_{*}(A,M)\cong\operatorname{H}\nolimits^{*}(\operatorname{\mathbf{T}}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits^{\prime}\otimes_{A}M).
Proof.

First, we prove (1)(1). Since 𝚺n𝐓\boldsymbol{\Sigma}^{n}\operatorname{\mathbf{T}}\nolimits is an acyclic complex of projective AA-bimodules for all nn\in\operatorname{\mathbb{Z}}\nolimits, it suffices to show that the induced morphism [𝐓,𝐓AM][\operatorname{\mathbf{T}}\nolimits,\operatorname{\mathbf{T}}\nolimits^{\prime}\otimes_{A}M] \rightarrow [𝐓,M][\operatorname{\mathbf{T}}\nolimits,M] is an isomorphism. Take a chain map u:𝐓Mu:\operatorname{\mathbf{T}}\nolimits\rightarrow M. Since εAidM\varepsilon^{\prime}\otimes_{A}\operatorname{id}\nolimits_{M} is an epimorphism and T0T_{0} is projective, there exists a morphism τ0:T0T0AM\tau_{0}:T_{0}\rightarrow T^{\prime}_{0}\otimes_{A}M such that (εAid)τ0=u(\varepsilon^{\prime}\otimes_{A}\operatorname{id}\nolimits)\tau_{0}=u. Then we get

(d0𝐓AidM)τ0d1𝐓\displaystyle(d_{0}^{\operatorname{\mathbf{T}}\nolimits^{\prime}}\otimes_{A}\operatorname{id}\nolimits_{M})\tau_{0}d^{\operatorname{\mathbf{T}}\nolimits}_{1} =(ηAidM)(εAidM)τ0d1𝐓\displaystyle=(\eta^{\prime}\otimes_{A}\operatorname{id}\nolimits_{M})(\varepsilon^{\prime}\otimes_{A}\operatorname{id}\nolimits_{M})\tau_{0}d^{\operatorname{\mathbf{T}}\nolimits}_{1}
=(ηAidM)ud1𝐓=0.\displaystyle=(\eta^{\prime}\otimes_{A}\operatorname{id}\nolimits_{M})ud^{\operatorname{\mathbf{T}}\nolimits}_{1}=0.

It follows from Lemma 3.4 that τ0\tau_{0} extends to a chain map τ:𝐓𝐓AM\tau:\operatorname{\mathbf{T}}\nolimits\rightarrow\operatorname{\mathbf{T}}\nolimits^{\prime}\otimes_{A}M such that 𝓂Ae(𝐓,εAidM)(τ)=u\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,\varepsilon^{\prime}\otimes_{A}\operatorname{id}\nolimits_{M})(\tau)=u. On the other hand, assume that τ:𝐓𝐓AM\tau:\operatorname{\mathbf{T}}\nolimits\rightarrow\operatorname{\mathbf{T}}\nolimits^{\prime}\otimes_{A}M is a chain map such that (εAid)τ:𝐓M(\varepsilon^{\prime}\otimes_{A}\operatorname{id}\nolimits)\tau:\operatorname{\mathbf{T}}\nolimits\rightarrow M is null-homotopic. Then there exists a morphism s:T1Ms:T_{-1}\rightarrow M such that (εAid)τ0=sd0𝐓(\varepsilon^{\prime}\otimes_{A}\operatorname{id}\nolimits)\tau_{0}=sd^{\operatorname{\mathbf{T}}\nolimits}_{0}. Since εAid\varepsilon^{\prime}\otimes_{A}\operatorname{id}\nolimits is an epimorphism and T1T_{-1} is projective, there exists a lifting h1:T1T0AMh_{-1}:T_{-1}\rightarrow T^{\prime}_{0}\otimes_{A}M of ss such that (εAid)h1=s(\varepsilon^{\prime}\otimes_{A}\operatorname{id}\nolimits)h_{-1}=s, and we have

(d0𝐓Aid)h1d0𝐓\displaystyle(d_{0}^{\operatorname{\mathbf{T}}\nolimits^{\prime}}\otimes_{A}\operatorname{id}\nolimits)h_{-1}d^{\operatorname{\mathbf{T}}\nolimits}_{0} =(ηAid)(εAid)h1d0𝐓\displaystyle=(\eta^{\prime}\otimes_{A}\operatorname{id}\nolimits)(\varepsilon^{\prime}\otimes_{A}\operatorname{id}\nolimits)h_{-1}d^{\operatorname{\mathbf{T}}\nolimits}_{0}
=(ηAid)sd0𝐓\displaystyle=(\eta^{\prime}\otimes_{A}\operatorname{id}\nolimits)sd^{\operatorname{\mathbf{T}}\nolimits}_{0}
=(d0𝐓Aid)τ0.\displaystyle=(d_{0}^{\operatorname{\mathbf{T}}\nolimits^{\prime}}\otimes_{A}\operatorname{id}\nolimits)\tau_{0}.

As in the proof of Lemma 1.1, we can construct h0:T0T1AMh_{0}:T_{0}\rightarrow T^{\prime}_{1}\otimes_{A}M such that τ0=(d1𝐓Aid)h0+h1(d0𝐓Aid)\tau_{0}=(d_{1}^{\operatorname{\mathbf{T}}\nolimits^{\prime}}\otimes_{A}\operatorname{id}\nolimits)h_{0}+h_{-1}(d_{0}^{\operatorname{\mathbf{T}}\nolimits^{\prime}}\otimes_{A}\operatorname{id}\nolimits). Inductively, we obtain a family {hi}i1\{h_{i}\}_{i\geq-1} satisfying τi=(di+1𝐓Aid)hi+hi1(di𝐓Aid)\tau_{i}=(d_{i+1}^{\operatorname{\mathbf{T}}\nolimits^{\prime}}\otimes_{A}\operatorname{id}\nolimits)h_{i}+h_{i-1}(d_{i}^{\operatorname{\mathbf{T}}\nolimits^{\prime}}\otimes_{A}\operatorname{id}\nolimits) for i0i\geq 0. It follows from Lemma 1.2 that {hi}i1\{h_{i}\}_{i\geq-1} extends to a null-homotopy of τ\tau.

In order to prove (2), take the AeA^{\textrm{e}}-dual complex 𝐓\operatorname{\mathbf{T}}\nolimits^{\vee} of 𝐓\operatorname{\mathbf{T}}\nolimits. It is still an acyclic complex of finitely generated projective AA-bimodules. Thus, the second statement follows from the first statement and the fact that there exists an isomorphism

𝐓^AeC𝓂Ae(𝐓,C)\displaystyle\operatorname{\mathbf{T}}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A^{\textrm{e}}}C\cong\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits^{\vee},C)

for any chain complex CC over AeA^{\textrm{e}}. ∎

Let 𝐓\operatorname{\mathbf{T}}\nolimits be a complete resolution of AA and MM and NN AA-bimodules. We consider the following two chain maps: the first is the composition map

(3.5) 𝓂Ae(𝐓,M)𝓂Ae(𝐓,𝐓AN)𝓂Ae(𝐓,MAN)\displaystyle\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,M)\otimes\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,\operatorname{\mathbf{T}}\nolimits\otimes_{A}N)\rightarrow\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,M\otimes_{A}N)

defined by

uv(uAidN)vu\otimes v\mapsto(u\otimes_{A}\operatorname{id}\nolimits_{N})\circ v

for any homogeneous element uv𝓂Ae(𝐓,M)𝓂Ae(𝐓,𝐓AN)u\otimes v\in\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,M)\otimes\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,\operatorname{\mathbf{T}}\nolimits\otimes_{A}N), and the second is the chain map

(3.6) 𝓂Ae(𝐓,M)(𝐓^Ae(𝐓AN))𝐓Ae(MAN)\displaystyle\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,M)\otimes(\operatorname{\mathbf{T}}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits\otimes_{A}N))\rightarrow\operatorname{\mathbf{T}}\nolimits\otimes_{A^{\textrm{e}}}(M\otimes_{A}N)

given by

u(xAe(xAn))(1)|u||x|xAe(u(x)An)u\otimes(x\otimes_{A^{\textrm{e}}}(x^{\prime}\otimes_{A}n))\mapsto(-1)^{|u||x|}x\otimes_{A^{\textrm{e}}}(u(x^{\prime})\otimes_{A}n)

for any homogeneous element u(xAe(xAn))𝓂Ae(𝐓,M)(𝐓^Ae(𝐓Au\otimes(x\otimes_{A^{\textrm{e}}}(x^{\prime}\otimes_{A}n))\in\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,M)\otimes(\operatorname{\mathbf{T}}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits\otimes_{A} N))N)). By Proposition 3.14, these chain maps induce a well-defined operators, called composition products,

(3.7) H^r(A,M)H^s(A,N)H^r+s(A,MAN),\displaystyle\operatorname{\widehat{H}}\nolimits^{r}(A,M)\otimes\operatorname{\widehat{H}}\nolimits^{s}(A,N)\rightarrow\operatorname{\widehat{H}}\nolimits^{r+s}(A,M\otimes_{A}N),
(3.8) H^r(A,M)H^s(A,N)H^sr(A,MAN).\displaystyle\operatorname{\widehat{H}}\nolimits^{r}(A,M)\otimes\operatorname{\widehat{H}}\nolimits_{s}(A,N)\rightarrow\operatorname{\widehat{H}}\nolimits_{s-r}(A,M\otimes_{A}N).
Theorem 3.15.

The following statements hold.

  1. (1)

    The composition product (3.7)(\ref{eq:7}) agrees with the cup product (3.3)(\ref{eq:8}).

  2. (2)

    The composition product (3.8)(\ref{eq:12}) agrees with the cap product (3.4)(\ref{eq:10}).

Proof.

First of all, we show (1)(1). The composition product (3.7)(\ref{eq:7}) is induced by the chain map (3.5)(\ref{eq:9}) via the quasi-isomorphism α:=𝓂Ae(𝐓,εAidM)\alpha:=\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,\varepsilon^{\prime}\otimes_{A}\operatorname{id}\nolimits_{M}). We now construct a quasi-inverse of α\alpha, i.e., a chain map

α:𝓂Ae(𝐓,N)𝓂Ae(𝐓,𝐓AN)\displaystyle\alpha^{\prime}:\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,N)\rightarrow\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,\operatorname{\mathbf{T}}\nolimits^{\prime}\otimes_{A}N)

inducing the inverse of the isomorphism H(α)\operatorname{H}\nolimits(\alpha). Let α\alpha^{\prime} be the chain map

𝓂Ae(𝐓,N)𝓂Ae(𝐓,𝐓AN)\displaystyle\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,N)\rightarrow\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,\operatorname{\mathbf{T}}\nolimits^{\prime}\otimes_{A}N)

determined by u(id𝐓^Au)Δ^u\mapsto(\operatorname{id}\nolimits_{\operatorname{\mathbf{T}}\nolimits}\operatorname{\widehat{\otimes}}\nolimits_{A}u)\widehat{\Delta}, where Δ^:𝐓𝐓^A𝐓\widehat{\Delta}:\operatorname{\mathbf{T}}\nolimits\rightarrow\operatorname{\mathbf{T}}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A}\operatorname{\mathbf{T}}\nolimits is a diagonal approximation. Then the composition αα:𝓂Ae(𝐓,N)𝓂Ae(𝐓,N)\alpha\alpha^{\prime}:\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,N)\rightarrow\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,N) is given by

αα=(εAN)(id𝐓^Au)Δ^=(ε^Au)Δ^=εu.\alpha\alpha^{\prime}=(\varepsilon\otimes_{A}N)(\operatorname{id}\nolimits_{\operatorname{\mathbf{T}}\nolimits}\operatorname{\widehat{\otimes}}\nolimits_{A}u)\widehat{\Delta}=(\varepsilon\operatorname{\widehat{\otimes}}\nolimits_{A}u)\widehat{\Delta}=\varepsilon\smile u.

Since [ε]=1H^0(A,A)[\varepsilon]=1\in\operatorname{\widehat{H}}\nolimits^{0}(A,A), we see that α\alpha^{\prime} is a quasi-inverse of α\alpha. We are now able to compare the composition product (3.7)(\ref{eq:7}) with the cup product (3.3)(\ref{eq:8}) as follows: for any u𝓂Ae(𝐓,M)u\in\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,M) and v𝓂Ae(𝐓,N)v\in\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,N),

uv\displaystyle u\otimes v uα(v)\displaystyle\mapsto u\otimes\alpha^{\prime}(v)
(u^Aid)(id^Av)Δ^\displaystyle\mapsto(u\operatorname{\widehat{\otimes}}\nolimits_{A}\operatorname{id}\nolimits)(\operatorname{id}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A}v)\widehat{\Delta}
=(u^Av)Δ^=uv.\displaystyle=(u\operatorname{\widehat{\otimes}}\nolimits_{A}v)\widehat{\Delta}=u\smile v.

Secondly, we prove (2). In view of the proof of (1)(1), we construct a weak-inverse β\beta^{\prime} of β:=id𝐓^Ae(εAidN):𝐓^Ae(𝐓AN)𝐓AeN\beta:=\operatorname{id}\nolimits_{\operatorname{\mathbf{T}}\nolimits}\operatorname{\widehat{\otimes}}\nolimits_{A^{\textrm{e}}}(\varepsilon\otimes_{A}\operatorname{id}\nolimits_{N}):\operatorname{\mathbf{T}}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits\otimes_{A}N)\rightarrow\operatorname{\mathbf{T}}\nolimits\otimes_{A^{\textrm{e}}}N. Let β\beta^{\prime} be the composition

𝐓AeNΔ^id(𝐓^A𝐓)AeN𝐓^Ae(𝐓AeN).\displaystyle\operatorname{\mathbf{T}}\nolimits\otimes_{A^{\textrm{e}}}N\xrightarrow{\widehat{\Delta}\otimes\operatorname{id}\nolimits}(\operatorname{\mathbf{T}}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A}\operatorname{\mathbf{T}}\nolimits)\otimes_{A^{\textrm{e}}}N\cong\operatorname{\mathbf{T}}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits\otimes_{A^{\textrm{e}}}N).

Note that there exists a commutative square of chain complexes

(𝐓^A𝐓)AeN\textstyle{(\operatorname{\mathbf{T}}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A}\operatorname{\mathbf{T}}\nolimits)\otimes_{A^{\textrm{e}}}N\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(id^Aε)Aeid\scriptstyle{(\operatorname{id}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A}\varepsilon)\otimes_{A^{\textrm{e}}}\operatorname{id}\nolimits}\scriptstyle{\cong}(𝐓AA)AeN\textstyle{(\operatorname{\mathbf{T}}\nolimits\otimes_{A}A)\otimes_{A^{\textrm{e}}}N\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}𝐓^A(𝐓AeN)\textstyle{\operatorname{\mathbf{T}}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A}(\operatorname{\mathbf{T}}\nolimits\otimes_{A^{\textrm{e}}}N)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id^A(εAeid)\scriptstyle{\operatorname{id}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A}(\varepsilon\otimes_{A^{\textrm{e}}}\operatorname{id}\nolimits)}𝐓A(AAeN)\textstyle{\operatorname{\mathbf{T}}\nolimits\otimes_{A}(A\otimes_{A^{\textrm{e}}}N)}

Thus we see that the morphism

ββ:𝐓AeN𝐓AeN\displaystyle\beta\beta^{\prime}:\operatorname{\mathbf{T}}\nolimits\otimes_{A^{\textrm{e}}}N\rightarrow\operatorname{\mathbf{T}}\nolimits\otimes_{A^{\textrm{e}}}N

is of the form σAeidN\sigma\otimes_{A^{\textrm{e}}}\operatorname{id}\nolimits_{N}, where σ:𝐓𝐓\sigma:\operatorname{\mathbf{T}}\nolimits\rightarrow\operatorname{\mathbf{T}}\nolimits is an augmentation-preserving chain map and hence is homotopic to id𝐓\operatorname{id}\nolimits_{\operatorname{\mathbf{T}}\nolimits}. It is easy to check that the two chain map (3.8)(\ref{eq:12}) and (3.4)(\ref{eq:10}) coincide via the weak-inverse β\beta^{\prime} on the chain level. This completes the proof. ∎

Remark 3.16.

In the proof of Theorem 3.15, we can directly construct the inverse of H(α)\operatorname{H}\nolimits(\alpha) when N=AN=A in the following way: let nn\in\operatorname{\mathbb{Z}}\nolimits and g𝓂Ae(𝐓,A)ng\in\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,A)^{n} a cocycle. Then there uniquely (up to homotopy) exists a cocycle g¯𝓂Ae(𝐓,𝐓)n\overline{g}\in\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,\operatorname{\mathbf{T}}\nolimits)^{n}. Indeed, since gg is a cocycle, there uniquely exists a morphism g:Cokerd1𝚺n𝐓T1g^{\prime}:\operatorname{Coker}\nolimits d^{\boldsymbol{\Sigma}^{-n}\operatorname{\mathbf{T}}\nolimits}_{1}\rightarrow T_{-1} making the center square in the following diagram commute:

Cokerd1𝚺n𝐓\textstyle{\operatorname{Coker}\nolimits d^{\boldsymbol{\Sigma}^{-n}\operatorname{\mathbf{T}}\nolimits}_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}g\scriptstyle{g^{\prime}}Tn+1\textstyle{\cdots\rightarrow T_{n+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d1𝚺n𝐓\scriptstyle{d^{\boldsymbol{\Sigma}^{-n}\operatorname{\mathbf{T}}\nolimits}_{1}}Tn\textstyle{T_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d0𝚺n𝐓\scriptstyle{d^{\boldsymbol{\Sigma}^{-n}\operatorname{\mathbf{T}}\nolimits}_{0}}g\scriptstyle{g}Tn1\textstyle{T_{n-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d1𝚺n𝐓\scriptstyle{d^{\boldsymbol{\Sigma}^{-n}\operatorname{\mathbf{T}}\nolimits}_{-1}}Tn2\textstyle{T_{n-2}\rightarrow\cdots}T1\textstyle{\cdots\rightarrow T_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d1\scriptstyle{d_{1}}T0\textstyle{T_{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d0\scriptstyle{d_{0}}T1\textstyle{T_{-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d1\scriptstyle{d_{-1}}T2\textstyle{T_{-2}\rightarrow\cdots}A\textstyle{A\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

By Lemmas 1.1 and 1.2, there uniquely (up to homotopy) exists a lifting chain map g¯:𝚺n𝐓𝐓\overline{g}:\boldsymbol{\Sigma}^{-n}\operatorname{\mathbf{T}}\nolimits\rightarrow\operatorname{\mathbf{T}}\nolimits of gg. Moreover, we can take dn𝓂Ae(𝐓,𝐓)(g¯)d^{\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,\operatorname{\mathbf{T}}\nolimits)}_{n}(\overline{g}) as a lifting chain map of a coboundary dn𝓂Ae(𝐓,A)(g)d^{\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,A)}_{n}(g). Thus, we obtain the morphism

β:H^n(A,A)Hn(𝓂Ae(𝐓,𝐓))\displaystyle\beta:\operatorname{\widehat{H}}\nolimits^{n}(A,A)\rightarrow\operatorname{H}\nolimits^{n}(\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,\operatorname{\mathbf{T}}\nolimits))

given by [g][g¯][g]\mapsto[\overline{g}]. This map is the inverse of H(α)\operatorname{H}\nolimits(\alpha). Indeed, we have

H(α)β([g])=H(α)([g¯])=[α(g¯)]=[g]\displaystyle\operatorname{H}\nolimits(\alpha)\beta([g])=\operatorname{H}\nolimits(\alpha)([\overline{g}])=[\alpha(\overline{g})]=[g]

for all [g]H^n(A,A)[g]\in\operatorname{\widehat{H}}\nolimits^{n}(A,A). Thus, we get β=H(α)\beta=\operatorname{H}\nolimits(\alpha^{\prime}). In conclusion, we can compute the cup product \smile via chain maps when N=AN=A.

As a corollary of Theorem 3.15, we have the following property with respect to the cup product and the cap product.

Corollary 3.17.

For all αH^(A,L),βH^(A,M)\alpha\in\operatorname{\widehat{H}}\nolimits^{*}(A,L),\beta\in\operatorname{\widehat{H}}\nolimits^{*}(A,M) and ωH^(A,N)\omega\in\operatorname{\widehat{H}}\nolimits_{*}(A,N), we have

(αβ)ω=α(βω).(\alpha\smile\beta)\frown\omega=\alpha\frown(\beta\frown\omega).
Proof.

In view of Theorem 3.15, it suffices to show the statement for the elements represented by u𝓂Ae(𝐓,L)u\in\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,L), v𝓂Ae(𝐓,𝐓AM)v\in\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,\operatorname{\mathbf{T}}\nolimits\otimes_{A}M) and z=xAe(yAn)𝐓^Ae(𝐓AN)z=x\otimes_{A^{\textrm{e}}}(y\otimes_{A}n)\in\operatorname{\mathbf{T}}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits\otimes_{A}N). Then we have, on the chain level,

(uv)z\displaystyle(u\smile v)\frown z =(1)(|u|+|v|)|x|xAe(uv)(y)An\displaystyle=(-1)^{(|u|+|v|)|x|}x\otimes_{A^{\textrm{e}}}(u\smile v)(y)\otimes_{A}n
=(1)(|u|+|v|)|x|xAe(uAid)v(y)An,\displaystyle=(-1)^{(|u|+|v|)|x|}x\otimes_{A^{\textrm{e}}}(u\otimes_{A}\operatorname{id}\nolimits)v(y)\otimes_{A}n,
u(vz)\displaystyle u\frown(v\frown z) =u((1)|v||x|xAev(y)An)\displaystyle=u\frown((-1)^{|v||x|}x\otimes_{A^{\textrm{e}}}v(y)\otimes_{A}n)
=(1)|u||x|+|v||x|xAe(uAid)v(y)An.\displaystyle=(-1)^{|u||x|+|v||x|}x\otimes_{A^{\textrm{e}}}(u\otimes_{A}\operatorname{id}\nolimits)v(y)\otimes_{A}n.

This finishes the proof. ∎

3.3. Comparison with singular Hochschild cohomology ring

Our aim in this subsection is to prove our main theorem. For this, we first recall the result of Eu and Schedler. Let fHom¯Ae(ΩAeiA,A)f\in\underline{\operatorname{Hom}\nolimits}_{A^{\textrm{e}}}(\Omega_{A^{\textrm{e}}}^{i}A,A) and gHom¯Ae(ΩAejA,A)g\in\underline{\operatorname{Hom}\nolimits}_{A^{\textrm{e}}}(\Omega_{A^{\textrm{e}}}^{j}A,A) with i,ji,j\in\operatorname{\mathbb{Z}}\nolimits. We define a morphism fgf\cup g by

fg:=fΩAei(g)Hom¯Ae(ΩAei+jA,A),\displaystyle f\cup g:=f\Omega_{A^{\textrm{e}}}^{i}(g)\in\underline{\operatorname{Hom}\nolimits}_{A^{\textrm{e}}}(\Omega_{A^{\textrm{e}}}^{i+j}A,A),

where ΩAei(g)Hom¯Ae(ΩAei+jA,ΩAeiA)\Omega_{A^{\textrm{e}}}^{i}(g)\in\underline{\operatorname{Hom}\nolimits}_{A^{\textrm{e}}}(\Omega_{A^{\textrm{e}}}^{i+j}A,\Omega_{A^{\textrm{e}}}^{i}A) is induced by an autoequivalence ΩAe:Ae-mod¯Ae-mod¯\Omega_{A^{\textrm{e}}}:A^{\textrm{e}}\textrm{-}\underline{\mathrm{mod}}\rightarrow A^{\textrm{e}}\textrm{-}\underline{\mathrm{mod}}. On the other hand, it follows from [ES09, Proposition 2.1.8] that there exists an isomorphism

(3.9) ΩAeiL¯AeNL¯AeΩAeiN\displaystyle\Omega_{A^{\textrm{e}}}^{i}L\underline{\otimes}_{A^{\textrm{e}}}N\cong L\underline{\otimes}_{A^{\textrm{e}}}\Omega_{A^{\textrm{e}}}^{i}N

for any finitely generated AA-bimodules LL and NN and ii\in\operatorname{\mathbb{Z}}\nolimits. Using this isomorphism, we define

:Hom¯Ae(ΩAeiA,A)(ΩAejA¯AeM)ΩAejiA¯AeM\displaystyle\cap:\underline{\operatorname{Hom}\nolimits}_{A^{\textrm{e}}}(\Omega_{A^{\textrm{e}}}^{i}A,A)\otimes(\Omega_{A^{\textrm{e}}}^{j}A\underline{\otimes}_{A^{\textrm{e}}}M)\rightarrow\Omega_{A^{\textrm{e}}}^{j-i}A\underline{\otimes}_{A^{\textrm{e}}}M

to be the morphism making the following square commute:

Hom¯Ae(ΩAeiA,A)(ΩAejA¯AeM)\textstyle{\underline{\operatorname{Hom}\nolimits}_{A^{\textrm{e}}}(\Omega_{A^{\textrm{e}}}^{i}A,A)\otimes(\Omega_{A^{\textrm{e}}}^{j}A\underline{\otimes}_{A^{\textrm{e}}}M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cap}\scriptstyle{\cong}ΩAejiA¯AeM\textstyle{\Omega_{A^{\textrm{e}}}^{j-i}A\underline{\otimes}_{A^{\textrm{e}}}M}Hom¯Ae(ΩAeiA,A)(ΩAeiA¯AeΩAejiM)\textstyle{\underline{\operatorname{Hom}\nolimits}_{A^{\textrm{e}}}(\Omega_{A^{\textrm{e}}}^{i}A,A)\otimes(\Omega_{A^{\textrm{e}}}^{i}A\underline{\otimes}_{A^{\textrm{e}}}\Omega_{A^{\textrm{e}}}^{j-i}M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ϕ\scriptstyle{\phi}A¯AeΩAejiM\textstyle{A\underline{\otimes}_{A^{\textrm{e}}}\Omega_{A^{\textrm{e}}}^{j-i}M\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}

where the morphism ϕ\phi is given by

[f](aAem)f(a)Aem.[f]\otimes(a\otimes_{A^{\textrm{e}}}m)\mapsto f(a)\otimes_{A^{\textrm{e}}}m.
Theorem 3.18 ([ES09, Theorem 2.1.15]).

We have the following statements.

  1. (1)

    The graded vector space

    Hom¯Ae(ΩAeA,A):=iHom¯Ae(ΩAeiA,A)\underline{\operatorname{Hom}\nolimits}_{A^{\textrm{e}}}(\Omega_{A^{\textrm{e}}}^{\bullet}A,A):=\bigoplus_{i\in\operatorname{\mathbb{Z}}\nolimits}\underline{\operatorname{Hom}\nolimits}_{A^{\textrm{e}}}(\Omega_{A^{\textrm{e}}}^{i}A,A)

    equipped with \cup forms a graded commutative algebra, which extends the cup product on H1(A,A)\operatorname{H}\nolimits^{\geq 1}(A,A).

  2. (2)

    The morphism \cap extends the cap product between H(A,A)\operatorname{H}\nolimits^{*}(A,A) and H(A,M)\operatorname{H}\nolimits_{*}(A,M) for a finitely generated AA-bimodule MM and satisfies the relation

    (fg)z=f(gz)(f\cup g)\cap z=f\cap(g\cap z)

    for any zΩAeA¯AeM,fHom¯Ae(ΩAeA,A)z\in\Omega_{A^{\textrm{e}}}^{*}A\underline{\otimes}_{A^{\textrm{e}}}M,f\in\underline{\operatorname{Hom}\nolimits}_{A^{\textrm{e}}}(\Omega_{A^{\textrm{e}}}^{*}A,A) and gHom¯Ae(ΩAeA,A)g\in\underline{\operatorname{Hom}\nolimits}_{A^{\textrm{e}}}(\Omega_{A^{\textrm{e}}}^{*}A,A).

It follows from Proposition 2.7 that there exist isomorphisms

H^(A,M)Hom¯Ae(ΩAeA,M)andH^(A,M)ΩAeA¯AeM\displaystyle\operatorname{\widehat{H}}\nolimits^{*}(A,M)\cong\underline{\operatorname{Hom}\nolimits}_{A^{\textrm{e}}}(\Omega_{A^{\textrm{e}}}^{*}A,M)\ \ \mbox{and}\ \ \operatorname{\widehat{H}}\nolimits_{*}(A,M)\cong\Omega_{A^{\textrm{e}}}^{*}A\underline{\otimes}_{A^{\textrm{e}}}M

for a finitely generated AA-bimodule MM. We will prove that \smile and \frown are equivalent to \cup and \cap, respectively, via the isomorphisms above.

Lemma 3.19.

We have the following statements.

  1. (1)

    There exists an isomorphism

    H^(A,A)Hom¯Ae(ΩAeA,A)\operatorname{\widehat{H}}\nolimits^{\bullet}(A,A)\cong\underline{\operatorname{Hom}\nolimits}_{A^{\textrm{e}}}(\Omega_{A^{\textrm{e}}}^{\bullet}A,A)

    of graded commutative algebras.

  2. (2)

    For any i,ji,j\in\operatorname{\mathbb{Z}}\nolimits, there exists a commutative diagram

    Hom¯Ae(ΩAeiA,A)(ΩAejA¯AeM)\textstyle{\underline{\operatorname{Hom}\nolimits}_{A^{\textrm{e}}}(\Omega_{A^{\textrm{e}}}^{i}A,A)\otimes(\Omega_{A^{\textrm{e}}}^{j}A\underline{\otimes}_{A^{\textrm{e}}}M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cap}ΩAejiA¯AeM\textstyle{\Omega_{A^{\textrm{e}}}^{j-i}A\underline{\otimes}_{A^{\textrm{e}}}M\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H^i(A,A)H^j(A,M)\textstyle{\operatorname{\widehat{H}}\nolimits^{i}(A,A)\otimes\operatorname{\widehat{H}}\nolimits_{j}(A,M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\frown}H^ji(A,M)\textstyle{\operatorname{\widehat{H}}\nolimits_{j-i}(A,M)}

    where MM is a finitely generated AA-bimodule and the vertical morphisms are given by the isomorphisms as in Proposition 2.7.

Proof.

In order to prove (1), we will show that the following diagram commutes for every i,ji,j\in\operatorname{\mathbb{Z}}\nolimits:

Hom¯Ae(ΩAeiA,A)Hom¯Ae(ΩAejA,A)\textstyle{\underline{\operatorname{Hom}\nolimits}_{A^{\textrm{e}}}(\Omega_{A^{\textrm{e}}}^{i}A,A)\otimes\underline{\operatorname{Hom}\nolimits}_{A^{\textrm{e}}}(\Omega_{A^{\textrm{e}}}^{j}A,A)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cup}ΦiΦj\scriptstyle{\Phi_{i}\otimes\Phi_{j}}Hom¯Ae(ΩAei+jA,A)\textstyle{\underline{\operatorname{Hom}\nolimits}_{A^{\textrm{e}}}(\Omega_{A^{\textrm{e}}}^{i+j}A,A)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Φi+j\scriptstyle{\Phi_{i+j}}H^i(A,A)H^j(A,A)\textstyle{\operatorname{\widehat{H}}\nolimits^{i}(A,A)\otimes\operatorname{\widehat{H}}\nolimits^{j}(A,A)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}idβ\scriptstyle{\operatorname{id}\nolimits\otimes\beta}H^i(A,A)Hj(𝓂Ae(𝐓A,𝐓A))\textstyle{\operatorname{\widehat{H}}\nolimits^{i}(A,A)\otimes\operatorname{H}\nolimits^{j}(\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits^{A},\operatorname{\mathbf{T}}\nolimits^{A}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H^i+j(A,M)\textstyle{\operatorname{\widehat{H}}\nolimits^{i+j}(A,M)}

where 𝐓A𝜀A\operatorname{\mathbf{T}}\nolimits^{A}\xrightarrow{\varepsilon}A is a minimal complete resolution, the lower horizontal morphism is the composition product, the morphism Φ\Phi_{*} is the isomorphism appeared in the proof of Proposition 2.7 and the morphism β\beta is the isomorphism constructed in Remark 3.16. Recall that we decompose each differential di𝐓Ad_{i}^{\operatorname{\mathbf{T}}\nolimits^{A}} as di𝐓A=ιiπid_{i}^{\operatorname{\mathbf{T}}\nolimits^{A}}=\iota_{i}\pi_{i} with πi:TiAΩAeiA\pi_{i}:T_{i}^{A}\rightarrow\Omega_{A^{\textrm{e}}}^{i}A and ιi:ΩAeiATi1A\iota_{i}:\Omega_{A^{\textrm{e}}}^{i}A\rightarrow T_{i-1}^{A}. Let [f][g]Hom¯Ae(ΩAeiA,A)Hom¯Ae(ΩAejA,A)[f]\otimes[g]\in\underline{\operatorname{Hom}\nolimits}_{A^{\textrm{e}}}(\Omega_{A^{\textrm{e}}}^{i}A,A)\otimes\underline{\operatorname{Hom}\nolimits}_{A^{\textrm{e}}}(\Omega_{A^{\textrm{e}}}^{j}A,A) be arbitrary. Recalling the definitions of β\beta and of the autoequivalence ΩAe\Omega_{A^{\textrm{e}}} of Ae-mod¯A^{\textrm{e}}\textrm{-}\underline{\mathrm{mod}}, we have the following commutative diagram with exact rows:

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ΩAel+j+1A\textstyle{\Omega_{A^{\textrm{e}}}^{l+j+1}A\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(1)jιl+j+1\scriptstyle{(-1)^{j}\iota_{l+j+1}}ΩAel+1(g)\scriptstyle{\Omega_{A^{\textrm{e}}}^{l+1}(g)}Tl+jA\textstyle{T_{l+j}^{A}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}πl+j\scriptstyle{\pi_{l+j}}gl\scriptstyle{g_{l}}ΩAel+jA\textstyle{\Omega_{A^{\textrm{e}}}^{l+j}A\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ΩAel(g)\scriptstyle{\Omega_{A^{\textrm{e}}}^{l}(g)}0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ΩAel+1A\textstyle{\Omega_{A^{\textrm{e}}}^{l+1}A\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ιl+1\scriptstyle{\iota_{l+1}}TlA\textstyle{T_{l}^{A}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}πl\scriptstyle{\pi_{l}}ΩAelA\textstyle{\Omega_{A^{\textrm{e}}}^{l}A\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

where ll\in\operatorname{\mathbb{Z}}\nolimits and the morphism glg_{l} is a ll-th component of a lifting cahin map of gπjg\pi_{j}. Then we have

[f][g]\displaystyle[f]\otimes[g] ΦiΦj[fπi][gπj]\displaystyle\stackrel{{\scriptstyle{\small\Phi_{i}\otimes\Phi_{j}}}}{{\longmapsto}}[f\pi_{i}]\otimes[g\pi_{j}]
idβ[fπi]β([gπj])\displaystyle\stackrel{{\scriptstyle\operatorname{id}\nolimits\otimes\beta}}{{\longmapsto}}[f\pi_{i}]\otimes\beta([g\pi_{j}])
[(fπi)gi]\displaystyle\longmapsto[(f\pi_{i})g_{i}]

and

[f][g]\displaystyle[f]\otimes[g] [fΩAei(g)]\displaystyle\stackrel{{\scriptstyle\cup}}{{\longmapsto}}[f\Omega_{A^{\textrm{e}}}^{i}(g)]
Φi+j[(fΩAei(g))πi+j]=[f(ΩAei(g)πi+j)]=[f(πigi)].\displaystyle\stackrel{{\scriptstyle\Phi_{i+j}}}{{\longmapsto}}[(f\Omega_{A^{\textrm{e}}}^{i}(g))\pi_{i+j}]=[f(\Omega_{A^{\textrm{e}}}^{i}(g)\pi_{i+j})]=[f(\pi_{i}g_{i})].

For the second statement, let 𝐓M\operatorname{\mathbf{T}}\nolimits^{M} be a minimal complete resolution of MM. For any ii\in\operatorname{\mathbb{Z}}\nolimits, there exist two exact sequences

0ΩAei+1Mιi+1TiMπiΩAei+1M0,\displaystyle 0\rightarrow\Omega_{A^{\textrm{e}}}^{i+1}M\xrightarrow{\iota_{i+1}}T_{i}^{M}\xrightarrow{\pi_{i}}\Omega_{A^{\textrm{e}}}^{i+1}M\rightarrow 0,
0AAΩAei+1MidAιi+1AATiMidAπiAAΩAei+1M0.\displaystyle 0\rightarrow A\otimes_{A}\Omega_{A^{\textrm{e}}}^{i+1}M\xrightarrow{\operatorname{id}\nolimits\otimes_{A}\iota_{i+1}}A\otimes_{A}T_{i}^{M}\xrightarrow{\operatorname{id}\nolimits\otimes_{A}\pi_{i}}A\otimes_{A}\Omega_{A^{\textrm{e}}}^{i+1}M\rightarrow 0.

It follows from [CJ14, Lemma 2.7] that H^r(A,TsM)=0\operatorname{\widehat{H}}\nolimits_{r}(A,T_{s}^{M})=0 for all r,sr,s\in\operatorname{\mathbb{Z}}\nolimits. Thus the property (QII2)({\rm QII}_{2}) of the cap product \frown in Proposition 3.12 implies that there exists a commutative square

H^i(A,A)H^i(A,ΩAejiM)\textstyle{\operatorname{\widehat{H}}\nolimits^{i}(A,A)\otimes\operatorname{\widehat{H}}\nolimits_{i}(A,\Omega_{A^{\textrm{e}}}^{j-i}M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\frown}H^0(A,ΩAejiM)\textstyle{\operatorname{\widehat{H}}\nolimits_{0}(A,\Omega_{A^{\textrm{e}}}^{j-i}M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H^i(A,A)H^j(A,M)\textstyle{\operatorname{\widehat{H}}\nolimits^{i}(A,A)\otimes\operatorname{\widehat{H}}\nolimits_{j}(A,M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\frown}H^0(A,ΩAejiM)\textstyle{\operatorname{\widehat{H}}\nolimits_{0}(A,\Omega_{A^{\textrm{e}}}^{j-i}M)}

where i,ji,j\in\operatorname{\mathbb{Z}}\nolimits and the two vertical morphisms are isomorphisms. In order to complete the proof of (2), it suffices to show that a square

Hom¯Ae(ΩAeiA,A)(ΩAeiA¯AeΩAejiM)\textstyle{\underline{\operatorname{Hom}\nolimits}_{A^{\textrm{e}}}(\Omega_{A^{\textrm{e}}}^{i}A,A)\otimes(\Omega_{A^{\textrm{e}}}^{i}A\underline{\otimes}_{A^{\textrm{e}}}\Omega_{A^{\textrm{e}}}^{j-i}M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ϕ\scriptstyle{\phi}ΦiΨi\scriptstyle{\Phi_{i}\otimes\Psi_{i}}A¯AeΩAejiM\textstyle{A\underline{\otimes}_{A^{\textrm{e}}}\Omega_{A^{\textrm{e}}}^{j-i}M\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ψ0\scriptstyle{\Psi_{0}}H^i(A,A)H^i(A,ΩAejiM)\textstyle{\operatorname{\widehat{H}}\nolimits^{i}(A,A)\otimes\operatorname{\widehat{H}}\nolimits_{i}(A,\Omega_{A^{\textrm{e}}}^{j-i}M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\frown}H^0(A,ΩAejiM)\textstyle{\operatorname{\widehat{H}}\nolimits_{0}(A,\Omega_{A^{\textrm{e}}}^{j-i}M)}

is commutative for all i,ji,j\in\operatorname{\mathbb{Z}}\nolimits. Let

[f](xAem)Hom¯Ae(ΩAeiA,A)(ΩAeiA¯AeΩAejiM)[f]\otimes(x\otimes_{A^{\textrm{e}}}m)\in\underline{\operatorname{Hom}\nolimits}_{A^{\textrm{e}}}(\Omega_{A^{\textrm{e}}}^{i}A,A)\otimes(\Omega_{A^{\textrm{e}}}^{i}A\underline{\otimes}_{A^{\textrm{e}}}\Omega_{A^{\textrm{e}}}^{j-i}M)

be arbitrary. Since πiAeidΩAejiM\pi_{i}\otimes_{A^{\textrm{e}}}\operatorname{id}\nolimits_{\Omega_{A^{\textrm{e}}}^{j-i}M} is an epimorphism, we have xAem=(πiAeid)(aAem)x\otimes_{A^{\textrm{e}}}m=(\pi_{i}\otimes_{A^{\textrm{e}}}\operatorname{id}\nolimits)(a\otimes_{A^{\textrm{e}}}m) for some aTiAa\in T_{i}^{A}. Let Δ^:𝐓A𝐓A^A𝐓A\widehat{\Delta}:\operatorname{\mathbf{T}}\nolimits^{A}\rightarrow\operatorname{\mathbf{T}}\nolimits^{A}\operatorname{\widehat{\otimes}}\nolimits_{A}\operatorname{\mathbf{T}}\nolimits^{A} be a diagonal approximation, and we denote

Δ^(a)=(xilAyl)l(𝐓A^A𝐓A)i.\widehat{\Delta}(a)=(x_{i-l}\otimes_{A}y_{l})_{l\in\operatorname{\mathbb{Z}}\nolimits}\in(\operatorname{\mathbf{T}}\nolimits^{A}\operatorname{\widehat{\otimes}}\nolimits_{A}\operatorname{\mathbf{T}}\nolimits^{A})_{i}.

Then we have

[f](xAem)\displaystyle[f]\otimes(x\otimes_{A^{\textrm{e}}}m) ΦiΨi[fπi][aAem]\displaystyle\stackrel{{\scriptstyle\Phi_{i}\otimes\Psi_{i}}}{{\longmapsto}}[f\pi_{i}]\otimes[a\otimes_{A^{\textrm{e}}}m]
[x0Aefπi(yi)m].\displaystyle\stackrel{{\scriptstyle\frown}}{{\longmapsto}}[x_{0}\otimes_{A^{\textrm{e}}}f\pi_{i}(y_{i})m].

On the other hand, since εfπi\varepsilon\smile f\pi_{i} is homotopic to fπif\pi_{i}, there exists a morphism h:Ti1AAh:T_{i-1}^{A}\rightarrow A such that (εfπi)fπi=hdi𝐓A(\varepsilon\smile f\pi_{i})-f\pi_{i}=hd_{i}^{\operatorname{\mathbf{T}}\nolimits^{A}}. Recalling the definition of the cup product \smile on H^\operatorname{\widehat{H}}\nolimits, we have

(εAeid)(x0fπi(yi)Aem)\displaystyle(\varepsilon\otimes_{A^{\textrm{e}}}\operatorname{id}\nolimits)(x_{0}f\pi_{i}(y_{i})\otimes_{A^{\textrm{e}}}m)
=\displaystyle= ε(x0)fπi(yi)Aem\displaystyle\ \varepsilon(x_{0})f\pi_{i}(y_{i})\otimes_{A^{\textrm{e}}}m
=\displaystyle= (εfπi)(a)Aem\displaystyle\ (\varepsilon\smile f\pi_{i})(a)\otimes_{A^{\textrm{e}}}m
=\displaystyle= fπi(a)Aem(hAeid)(di𝐓AAeid)(aAem)\displaystyle\ f\pi_{i}(a)\otimes_{A^{\textrm{e}}}m-(h\otimes_{A^{\textrm{e}}}\operatorname{id}\nolimits)(d_{i}^{\operatorname{\mathbf{T}}\nolimits^{A}}\otimes_{A^{\textrm{e}}}\operatorname{id}\nolimits)(a\otimes_{A^{\textrm{e}}}m)
=\displaystyle= f(x)Aem.\displaystyle\ f(x)\otimes_{A^{\textrm{e}}}m.

Therefore, we get

[f](xAem)\displaystyle[f]\otimes(x\otimes_{A^{\textrm{e}}}m) [f(x)Aem]\displaystyle\stackrel{{\scriptstyle\cap}}{{\longmapsto}}[f(x)\otimes_{A^{\textrm{e}}}m]
Ψ0[x0fπi(yi)Aem].\displaystyle\stackrel{{\scriptstyle\Psi_{0}}}{{\longmapsto}}[x_{0}f\pi_{i}(y_{i})\otimes_{A^{\textrm{e}}}m].

This completes the proof. ∎

We are now able to prove our main theorem. Wang in [Wan18] introduced singular Hochschild cochain complex Csg(A,A)C_{\textrm{sg}}(A,A) for any algebra AA over a field kk and defined the cup product sg\cup_{\rm sg} on Csg(A,A)C_{\textrm{sg}}(A,A). We now recall the definitions of Csg(A,A)C_{\textrm{sg}}(A,A) and of sg\cup_{\rm sg} in [Wan18]. The singular Hochschild cochain complex Csg(A,A)C_{\textrm{sg}}(A,A) of AA is defined by the inductive limit of the inductive system of Hochschild cochain complexes

C(A,A)θ0C(A,Ωnc(A))C(A,Ωncp(A))θpC(A,Ωncp+1(A)),\displaystyle C(A,A)\stackrel{{\scriptstyle\theta_{0}}}{{\hookrightarrow}}C(A,\Omega_{\rm nc}(A))\hookrightarrow\cdots\hookrightarrow C(A,\Omega_{\rm nc}^{p}(A))\stackrel{{\scriptstyle\theta_{p}}}{{\hookrightarrow}}C(A,\Omega_{\rm nc}^{p+1}(A))\hookrightarrow\cdots,

where C(A,Ωncp(A)):=𝓂Ae(Bar(A),Ωncp(A))C(A,\Omega_{\rm nc}^{p}(A)):=\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\mathrm{Bar}(A),\Omega_{\rm nc}^{p}(A)) with the AA-bimodule Ωncp(A):=AA¯p\Omega_{\rm nc}^{p}(A):=A\otimes\overline{A}^{\otimes p} concentrated in degree pp of which the left action is the multiplication of AA and the right action is defined by

(a0a¯1,p)ap+1:=i=0p(1)pia0a¯1,iaiai+1¯a¯i+2,p+1\displaystyle(a_{0}\otimes\overline{a}_{1,p})a_{p+1}:=\sum_{i=0}^{p}(-1)^{p-i}a_{0}\otimes\overline{a}_{1,i}\otimes\overline{a_{i}a_{i+1}}\otimes\overline{a}_{i+2,p+1}

for ap+1Aa_{p+1}\in A and a0a¯1,pΩncp(A)a_{0}\otimes\overline{a}_{1,p}\in\Omega_{\rm nc}^{p}(A), and the morphism θp\theta_{p} is defined as

C(A,Ωncp(A))C(A,Ωncp+1(A));ffidA¯.C(A,\Omega_{\rm nc}^{p}(A))\rightarrow C(A,\Omega_{\rm nc}^{p+1}(A));\quad f\mapsto f\otimes\operatorname{id}\nolimits_{\overline{A}}.

Here we have used the canonical isomorphism

HomAe(Bar(A)i,Ωncp(A))Homk(A¯i,Ωncp(A)).\operatorname{Hom}\nolimits_{A^{\textrm{e}}}(\mathrm{Bar}(A)_{i},\Omega_{\rm nc}^{p}(A))\cong\operatorname{Hom}\nolimits_{k}(\overline{A}^{\otimes i},\Omega_{\rm nc}^{p}(A)).

For ii\in\operatorname{\mathbb{Z}}\nolimits, we denote

HHsgi(A,A):=Hi(Csg(A,A)).\operatorname{HH}\nolimits_{\rm sg}^{i}(A,A):=\operatorname{H}\nolimits^{i}(C_{\textrm{sg}}(A,A)).

Moreover, for m,n,p,qm,n,p,q\in\operatorname{\mathbb{Z}}\nolimits, the cup product

sg:Cmp(A,Ωncp(A))Cnq(A,Ωncq(A))Cm+npq(A,Ωncp+q(A))\displaystyle\cup_{\rm sg}:C^{m-p}(A,\Omega_{\rm nc}^{p}(A))\otimes C^{n-q}(A,\Omega_{\rm nc}^{q}(A))\rightarrow C^{m+n-p-q}(A,\Omega_{\rm nc}^{p+q}(A))

is defined by

fg(μidA¯p+q)(idAfidA¯m)(gidA¯q),f\otimes g\mapsto\left(\mu\otimes\operatorname{id}\nolimits_{\overline{A}}^{\otimes p+q}\right)\left(\operatorname{id}\nolimits_{A}\otimes f\otimes\operatorname{id}\nolimits_{\overline{A}}^{\otimes m}\right)\left(g\otimes\operatorname{id}\nolimits_{\overline{A}}^{\otimes q}\right),

where μ:AAA\mu:A\otimes A\rightarrow A is the multiplication. Wang has proved the following two results.

Proposition 3.20 ([Wan18, Proposition 4.2 and Corollary 4.2]).

Under the same notation above, the singular Hochschild cochain complex Csg(A,A)C_{{\rm sg}}(A,A) equipped with the cup product sg\cup_{\rm sg} forms a differential graded associative algebra such that the induced cohomology ring HHsg(A,A)\operatorname{HH}\nolimits_{\rm sg}^{\bullet}(A,A) is graded commutative.

Before the second Wang’s result, we recall the definition of the singularity categories. Let AA be a (two-sided) Noetherian algebra AA over a field kk, and let 𝒟b(A)\mathcal{D}^{\textrm{b}}(A) be the bounded derived category of finitely generated AA-modules. Then the singularity category 𝒟sg(A)\mathcal{D}_{{\rm sg}}(A) of AA is defined to be the Verdier quotient 𝒟sg(A)=𝒟b(A)/𝒦b(projA)\mathcal{D}_{{\rm sg}}(A)=\mathcal{D}^{\textrm{b}}(A)/\mathcal{K}^{\textrm{b}}(\mathrm{proj}A), where 𝒦b(projA)\mathcal{K}^{\textrm{b}}(\mathrm{proj}A) is the bounded homotopy category of finitely generated projective AA-modules.

Proposition 3.21 ([Wan18, Proposition 4.7]).

Let AA be a Noetherian algebra over a field kk. Then there exists an isomorphism

HHsg(A,A)iHom𝒟sg(Ae)(A,𝚺iA)\operatorname{HH}\nolimits_{{\rm sg}}^{\bullet}(A,A)\rightarrow\bigoplus_{i\in\operatorname{\mathbb{Z}}\nolimits}\operatorname{Hom}\nolimits_{\mathcal{D}_{{\rm sg}}(A^{\textrm{e}})}(A,\boldsymbol{\Sigma}^{i}A)

of graded commutative algebras ((of degree 0)0), where the product on the right hand side is given by the Yoneda product.

If AA is a finite dimensional Frobenius algebra, then [Ric89, Theorem 2.1] implies that the canonical functor F:AeF:A^{\textrm{e}}-mod¯𝒟sg(Ae)\underline{\mathrm{mod}}\rightarrow\mathcal{D}_{\textrm{sg}}(A^{\textrm{e}}) is an equivalence of triangulated categories such that FΩAe𝚺1FF\circ\Omega_{A^{\textrm{e}}}\simeq\boldsymbol{\Sigma}^{-1}\circ F. Thus we have an isomorphism

Hom¯Ae(ΩAeA,A)iHom𝒟sg(Ae)(A,𝚺iA)\underline{\operatorname{Hom}\nolimits}_{A^{\textrm{e}}}(\Omega_{A^{\textrm{e}}}^{\bullet}A,A)\rightarrow\bigoplus_{i\in\operatorname{\mathbb{Z}}\nolimits}\operatorname{Hom}\nolimits_{\mathcal{D}_{\textrm{sg}}(A^{\textrm{e}})}(A,\boldsymbol{\Sigma}^{i}A)

of graded algebras. Consequently, from Lemma 3.19(1)\ref{lem:12}(1), we have the following result, which is our main theorem.

Theorem 3.22.

Let kk be a field and AA a finite dimensional Frobenius kk-algebra. Then there exists an isomorphism

H^(A,A)HHsg(A,A)\operatorname{\widehat{H}}\nolimits^{\bullet}(A,A)\cong\operatorname{HH}\nolimits_{{\rm sg}}^{\bullet}(A,A)

as graded commutative algebras.

It is easily checked that Tate-Hochschild cohomology rings are derived invariants of finite dimensional Frobenius algebras. Indeed, suppose that two finite dimensional Frobenius algebras AA and BB are derived equivalent, i.e., 𝒟b(A)\mathcal{D}^{{\rm b}}(A) is equivalent to 𝒟b(B)\mathcal{D}^{{\rm b}}(B) as triangulated categories. Then there exists an equivalence Fsg:𝒟sg(Ae)𝒟sg(Be)F_{\rm sg}:\mathcal{D}_{{\rm sg}}(A^{\textrm{e}})\rightarrow\mathcal{D}_{{\rm sg}}(B^{\rm e}) of triangulated categories such that Fsg(A)BF_{\rm sg}(A)\cong B (see [Zim14, Section 6] for instance). Then we see that the isomorphism

Hom𝒟sg(Ae)(A,𝚺iA)Hom𝒟sg(Be)(B,𝚺iB)\operatorname{Hom}\nolimits_{\mathcal{D}_{\textrm{sg}}(A^{\textrm{e}})}(A,\boldsymbol{\Sigma}^{i}A)\rightarrow\operatorname{Hom}\nolimits_{\mathcal{D}_{\textrm{sg}}(B^{\rm e})}(B,\boldsymbol{\Sigma}^{i}B)

induced by FsgF_{\rm sg} commutes with the Yoneda products. Consequently, Theorem 3.22 yields our claim.

4. Duality theorems in Tate-Hochschild theory

Let AA be a finite dimensional Frobenius algebra over a field kk. In this section, we prove that the Tate-Hochschild duality

H^n(A,M)H^n1(A,Mν11)\displaystyle\operatorname{\widehat{H}}\nolimits^{n}(A,M)\cong\operatorname{\widehat{H}}\nolimits_{-n-1}(A,{}_{1}M_{\nu^{-1}})

appeared in Section 2 is induced by the cap product for any integer nn and any AA-bimodule MM, and we prove that the cup product on Tate-Hochschild cohomology extends the cup product and the cap product on Hochschild (co)homology.

Applying the duality above for n=0n=0 and M=AM=A, we have

H^0(A,A)H^1(A,Aν11).\operatorname{\widehat{H}}\nolimits^{0}(A,A)\cong\operatorname{\widehat{H}}\nolimits_{-1}(A,{}_{1}A_{\nu^{-1}}).

Then an element ωH^1(A,Aν11)\omega\in\operatorname{\widehat{H}}\nolimits_{-1}(A,{}_{1}A_{\nu^{-1}}) is called the fundamental class of AA if the image under the isomorphism above of ω\omega is equal to 1H^0(A,A)1\in\operatorname{\widehat{H}}\nolimits^{0}(A,A).

Theorem 4.1.

The fundamental class ωH^1(A,Aν11)\omega\in\operatorname{\widehat{H}}\nolimits_{-1}(A,{}_{1}A_{\nu^{-1}}) induces an isomorphism – ω:H^n(A,M)H^n1(A,Mν11)\frown\omega:\operatorname{\widehat{H}}\nolimits^{n}(A,M)\rightarrow\operatorname{\widehat{H}}\nolimits_{-n-1}(A,{}_{1}M_{\nu^{-1}}) for any nn\in\operatorname{\mathbb{Z}}\nolimits and any AA-bimodule MM.

Proof.

If 𝐓\operatorname{\mathbf{T}}\nolimits is a complete resolution of AA, there exist isomorphisms

𝓂Ae(𝐓,M)𝐓AeM(𝐓)1ν1AeMν11,\displaystyle\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,M)\cong\operatorname{\mathbf{T}}\nolimits^{\vee}\otimes_{A^{\textrm{e}}}M\cong{}_{\nu^{-1}}(\operatorname{\mathbf{T}}\nolimits^{\vee})_{1}\otimes_{A^{\textrm{e}}}{}_{1}M_{\nu^{-1}},

where (𝐓)1ν1{}_{\nu^{-1}}(\operatorname{\mathbf{T}}\nolimits^{\vee})_{1} is an acyclic chain complex of finitely generated projective AA-bimodules. Observe that it is the 1-shifted complex of some complete resolution of the bimodule Aν1ν1{}_{\nu^{-1}}A_{\nu^{-1}}. Thus, there exists a complete resolution 𝐓\operatorname{\mathbf{T}}\nolimits^{\prime} of AA such that 𝚺𝐓(𝐓)1ν1\boldsymbol{\Sigma}\operatorname{\mathbf{T}}\nolimits^{\prime}\cong{}_{\nu^{-1}}(\operatorname{\mathbf{T}}\nolimits^{\vee})_{1}, so that we have isomorphisms

H^n(A,M)Hn(𝚺𝐓AeMν11)H^n1(A,Mν11).\displaystyle\operatorname{\widehat{H}}\nolimits^{n}(A,M)\cong\operatorname{H}\nolimits_{-n}(\boldsymbol{\Sigma}\operatorname{\mathbf{T}}\nolimits^{\prime}\otimes_{A^{\textrm{e}}}{}_{1}M_{\nu^{-1}})\cong\operatorname{\widehat{H}}\nolimits_{-n-1}(A,{}_{1}M_{\nu^{-1}}).

Clearly, the composite is natural in MM and compatible with long exact sequences in the following sense: for any short exact sequence of AA-bimodules

0LMN0,0\rightarrow L\rightarrow M\rightarrow N\rightarrow 0,

there exists a commutative diagram with long exact sequences

\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H^n(A,M)\textstyle{\operatorname{\widehat{H}}\nolimits^{n}(A,M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}H^n(A,N)\textstyle{\operatorname{\widehat{H}}\nolimits^{n}(A,N)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}H^n+1(A,L)\textstyle{\operatorname{\widehat{H}}\nolimits^{n+1}(A,L)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}\textstyle{\cdots}\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H^n1(A,Mν11)\textstyle{\operatorname{\widehat{H}}\nolimits_{-n-1}(A,{}_{1}M_{\nu^{-1}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H^n1(A,Nν11)\textstyle{\operatorname{\widehat{H}}\nolimits_{-n-1}(A,{}_{1}N_{\nu^{-1}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H^n2(A,Lν11)\textstyle{\operatorname{\widehat{H}}\nolimits_{-n-2}(A,{}_{1}L_{\nu^{-1}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\cdots}

Let φnM\varphi^{M}_{n} denote the isomorphism H^n(A,M)H^n1(A,Mν11)\operatorname{\widehat{H}}\nolimits^{n}(A,M)\rightarrow\operatorname{\widehat{H}}\nolimits_{-n-1}(A,{}_{1}M_{\nu^{-1}}). We claim that φ0M\varphi^{M}_{0} is given by the cap product with ωHH^1(A,Aν11)\omega\in\operatorname{\widehat{HH}}\nolimits_{-1}(A,{}_{1}A_{\nu^{-1}}). For any uH^0(A,M)u\in\operatorname{\widehat{H}}\nolimits^{0}(A,M), there exists a morphism f:AMf:A\rightarrow M of AA-bimodules such that uu is represented by ff and such that ff induces morphisms

H^0(A,A)H^0(A,M)andH^1(A,Aν11)H^1(A,Mν11)\displaystyle\operatorname{\widehat{H}}\nolimits^{0}(A,A)\rightarrow\operatorname{\widehat{H}}\nolimits^{0}(A,M)\quad\mbox{and}\quad\operatorname{\widehat{H}}\nolimits_{-1}(A,{}_{1}A_{\nu^{-1}})\rightarrow\operatorname{\widehat{H}}\nolimits_{-1}(A,{}_{1}M_{\nu^{-1}})

given by vuvv\mapsto u\smile v and by wuww\mapsto u\frown w, respectively. Hence the naturality of φ0\varphi^{*}_{0} implies that we have

φ0M([u])=φ0M([u]1)=[u]φ0A(1)=[u]ω.\displaystyle\varphi^{M}_{0}([u])=\varphi^{M}_{0}([u]\smile 1)=[u]\smile\varphi^{A}_{0}(1)=[u]\frown\omega.

A dimension-shifting argument shows that φM\varphi^{M}_{*} coincides with – ω\frown\omega in all degrees. ∎

It follows from Corollary 3.17 and Theorem 4.1 that the cup product is equivalent to the cap product in the following sense.

Corollary 4.2.

For any AA-bimodules M,NM,N and r,sr,s\in\operatorname{\mathbb{Z}}\nolimits, there exists a commutative diagram

H^r(A,M)H^s(A,N)\textstyle{\operatorname{\widehat{H}}\nolimits^{r}(A,M)\otimes\operatorname{\widehat{H}}\nolimits^{s}(A,N)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\smile}id(ω)\scriptstyle{\operatorname{id}\nolimits\otimes(-\frown\,\omega)}H^r+s(A,MAN)\textstyle{\operatorname{\widehat{H}}\nolimits^{r+s}(A,M\otimes_{A}N)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ω\scriptstyle{-\frown\,\omega}H^r(A,M)HH^s1(A,Nν11)\textstyle{\operatorname{\widehat{H}}\nolimits^{r}(A,M)\otimes\operatorname{\widehat{HH}}\nolimits_{-s-1}(A,{}_{1}N_{\nu^{-1}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\frown}HH^rs1(A,(MAN)ν11)\textstyle{\operatorname{\widehat{HH}}\nolimits_{-r-s-1}(A,{}_{1}(M\otimes_{A}N)_{\nu^{-1}})}

where the two vertical morphisms are isomorphisms.

Let νAut(A)\nu\in\operatorname{Aut}\nolimits(A) be the Nakayama automorphism of AA. It is known that there exist two Tate-Hochschild duality results for Frobenius algebras:

H^i(A,A)D(H^i1(A,A)) and H^i(A,A)D(H^i1(A,Aν21))\displaystyle\operatorname{\widehat{H}}\nolimits_{i}(A,A)\cong D(\operatorname{\widehat{H}}\nolimits_{-i-1}(A,A))\mbox{ and }\operatorname{\widehat{H}}\nolimits^{i}(A,A)\cong D(\operatorname{\widehat{H}}\nolimits^{-i-1}(A,{}_{1}A_{\nu^{2}}))

for any ii\in\operatorname{\mathbb{Z}}\nolimits, where the first is proved by Eu and Schedler in [ES09] and the second is proved by Bergh and Jorgensen in [BJ13]. We will give another proof of the two Tate-Hochschild duality results.

Corollary 4.3.

Let ν\nu be the Nakayama automorphism of AA. Then there exist two isomorphisms

H^i(A,A)D(H^i1(A,A)) and H^i(A,A)D(H^i1(A,Aν21))\displaystyle\operatorname{\widehat{H}}\nolimits_{i}(A,A)\cong D(\operatorname{\widehat{H}}\nolimits_{-i-1}(A,A))\mbox{ and }\operatorname{\widehat{H}}\nolimits^{i}(A,A)\cong D(\operatorname{\widehat{H}}\nolimits^{-i-1}(A,{}_{1}A_{\nu^{2}}))

for all ii\in\operatorname{\mathbb{Z}}\nolimits.

Proof.

Let 𝐓\operatorname{\mathbf{T}}\nolimits be a complete resolution of AA and MM a finitely generated AA-bimodule. The adjointness of 𝓂\operatorname{\operatorname{\mathscr{Hom}}\nolimits} and \otimes implies that there exists an isomorphism

𝓂Ae(𝐓,D(M))𝔻(𝐓AeM).\displaystyle\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,D(M))\cong\operatorname{\mathbb{D}}\nolimits(\operatorname{\mathbf{T}}\nolimits\otimes_{A^{\textrm{e}}}M).

Thus we have for any ii\in\operatorname{\mathbb{Z}}\nolimits

H^i(A,D(M))Hi(𝔻(𝐓AeM))D(H^i(A,M)).\displaystyle\operatorname{\widehat{H}}\nolimits^{i}(A,D(M))\cong\operatorname{H}\nolimits^{i}(\operatorname{\mathbb{D}}\nolimits(\operatorname{\mathbf{T}}\nolimits\otimes_{A^{\textrm{e}}}M))\cong D(\operatorname{\widehat{H}}\nolimits_{i}(A,M)).

If M=D(A)M=D(A), then we get

H^i(A,A)D(H^i(A,D(A)))D(H^i(A,Aν1))D(H^i1(A,Aν21)),\displaystyle\operatorname{\widehat{H}}\nolimits^{i}(A,A)\cong D(\operatorname{\widehat{H}}\nolimits_{i}(A,D(A)))\cong D(\operatorname{\widehat{H}}\nolimits_{i}(A,{}_{1}A_{\nu}))\cong D(\operatorname{\widehat{H}}\nolimits^{-i-1}(A,{}_{1}A_{\nu^{2}})),

where the second isomorphism is induced by the isomorphism D(A)Aν1D(A)\cong{}_{1}A_{\nu} of AA-bimodules, and the third isomorphism is induced by the isomorphism of Theorem 4.1. Similarly, if M=AM=A, then we obtain isomorphisms

D(H^i(A,A))H^i(A,D(A))H^i(A,Aν1)H^i1(A,A).\displaystyle D(\operatorname{\widehat{H}}\nolimits_{i}(A,A))\cong\operatorname{\widehat{H}}\nolimits^{i}(A,D(A))\cong\operatorname{\widehat{H}}\nolimits^{i}(A,{}_{1}A_{\nu})\cong\operatorname{\widehat{H}}\nolimits_{-i-1}(A,A).

Our next and last aim is to show that the cup product on Tate-Hochschild cohomology can be considered as an extension of the cup product and the cap product on Hochschild (co)homology.

Proposition 4.4.

Let M,NM,N be AA-bimodules and r,sr,s\in\operatorname{\mathbb{Z}}\nolimits. We denote by ^\widehat{\smile} the cup product on H^\operatorname{\widehat{H}}\nolimits and by ^\widehat{\frown} the cap product on H^\operatorname{\widehat{H}}\nolimits. Then the following two statements hold.

  1. (1)

    There exists a commutative square

    Hr(A,M)Hs(A,N)\textstyle{\operatorname{H}\nolimits^{r}(A,M)\otimes\operatorname{H}\nolimits^{s}(A,N)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\smile}Hr+s(A,MAN)\textstyle{\operatorname{H}\nolimits^{r+s}(A,M\otimes_{A}N)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H^r(A,M)H^s(A,N)\textstyle{\operatorname{\widehat{H}}\nolimits^{r}(A,M)\otimes\operatorname{\widehat{H}}\nolimits^{s}(A,N)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}^\scriptstyle{\widehat{\smile}}H^r+s(A,MAN)\textstyle{\operatorname{\widehat{H}}\nolimits^{r+s}(A,M\otimes_{A}N)}
  2. (2)

    There exist three commutative squares

    1. (a)

      the case r=0,s=0r=0,s=0

      H0(A,M)H^0(A,N)\textstyle{\operatorname{H}\nolimits^{0}(A,M)\otimes\operatorname{\widehat{H}}\nolimits_{0}(A,N)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\frown}H^0(A,MAN)\textstyle{\operatorname{\widehat{H}}\nolimits_{0}(A,M\otimes_{A}N)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H^0(A,M)H^0(A,N)\textstyle{\operatorname{\widehat{H}}\nolimits^{0}(A,M)\otimes\operatorname{\widehat{H}}\nolimits_{0}(A,N)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}^\scriptstyle{\widehat{\frown}}H^0(A,MAN)\textstyle{\operatorname{\widehat{H}}\nolimits_{0}(A,M\otimes_{A}N)}
    2. (b)

      the case r=0,s>0r=0,s>0

      H0(A,M)Hs(A,N)\textstyle{\operatorname{H}\nolimits^{0}(A,M)\otimes\operatorname{H}\nolimits_{s}(A,N)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\frown}Hs(A,MAN)\textstyle{\operatorname{H}\nolimits_{s}(A,M\otimes_{A}N)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H^0(A,M)H^s(A,N)\textstyle{\operatorname{\widehat{H}}\nolimits^{0}(A,M)\otimes\operatorname{\widehat{H}}\nolimits_{s}(A,N)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}^\scriptstyle{\widehat{\frown}}H^s(A,MAN)\textstyle{\operatorname{\widehat{H}}\nolimits_{s}(A,M\otimes_{A}N)}
    3. (c)

      the case r>0,s>0r>0,s>0 with srs\geq r

      Hr(A,M)Hs(A,N)\textstyle{\operatorname{H}\nolimits^{r}(A,M)\otimes\operatorname{H}\nolimits_{s}(A,N)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\frown}Hsr(A,MAN)\textstyle{\operatorname{H}\nolimits_{s-r}(A,M\otimes_{A}N)}H^r(A,M)H^s(A,N)\textstyle{\operatorname{\widehat{H}}\nolimits^{r}(A,M)\otimes\operatorname{\widehat{H}}\nolimits_{s}(A,N)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}^\scriptstyle{\widehat{\frown}}H^sr(A,MAN)\textstyle{\operatorname{\widehat{H}}\nolimits_{s-r}(A,M\otimes_{A}N)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

In all of the four diagrams above, the vertical morphisms consist of the morphisms constructed after Lemma 2.10.

Proof.

Let 𝐓\operatorname{\mathbf{T}}\nolimits be a complete resolution of AA, and we denote by 𝐏\operatorname{\mathbf{P}}\nolimits the truncation 𝐓0\operatorname{\mathbf{T}}\nolimits_{\geq 0}. Let Δ:𝐏𝐏A𝐏\Delta:\operatorname{\mathbf{P}}\nolimits\rightarrow\operatorname{\mathbf{P}}\nolimits\otimes_{A}\operatorname{\mathbf{P}}\nolimits be a diagonal approximation associated with a diagonal approximation Δ^:𝐓𝐓^A𝐓\widehat{\Delta}:\operatorname{\mathbf{T}}\nolimits\rightarrow\operatorname{\mathbf{T}}\nolimits\operatorname{\widehat{\otimes}}\nolimits_{A}\operatorname{\mathbf{T}}\nolimits, which has been constructed after Theorem 3.6.

Since the chain map Δ\Delta consists of non-negative components of Δ^=(Δ^p(n))\widehat{\Delta}=\prod(\prod\widehat{\Delta}^{(n)}_{p}), we have a commutative square

𝓂Ae(𝐏,M)𝓂Ae(𝐏,N)\textstyle{\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{P}}\nolimits,M)\otimes\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{P}}\nolimits,N)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\smile}𝓂Ae(ϑ,M)𝓂Ae(ϑ,N)\scriptstyle{\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\vartheta,M)\otimes\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\vartheta,N)}𝓂Ae(𝐏,MAN)\textstyle{\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{P}}\nolimits,M\otimes_{A}N)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝓂Ae(ϑ,MAN)\scriptstyle{\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\vartheta,M\otimes_{A}N)}𝓂Ae(𝐓,M)𝓂Ae(𝐓,N)\textstyle{\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,M)\otimes\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,N)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}^\scriptstyle{\widehat{\smile}}𝓂Ae(𝐓,MAN)\textstyle{\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,M\otimes_{A}N)}

where ϑ\vartheta is the canonical chain map 𝐓𝐏\operatorname{\mathbf{T}}\nolimits\rightarrow\operatorname{\mathbf{P}}\nolimits. Thus, we get the commutative square in (1)(1).

Recalling that H^0(A,N)H0(A,N)\operatorname{\widehat{H}}\nolimits_{0}(A,N)\leq\operatorname{H}\nolimits_{0}(A,N), we see that uzH^0(A,MAN)u\frown z\in\operatorname{\widehat{H}}\nolimits_{0}(A,M\otimes_{A}N) for uH0(A,M)u\in\operatorname{H}\nolimits^{0}(A,M) and zH^0(A,N)z\in\operatorname{\widehat{H}}\nolimits_{0}(A,N). Thus we obtain the commutative square in ((a)) of (2)(2).

For any r0r\geq 0 and s>0s>0 with srs\geq r, we have a commutative square

𝓂Ae(𝐓,M)r(𝐓AeM)s\textstyle{\operatorname{\operatorname{\mathscr{Hom}}\nolimits}_{A^{\textrm{e}}}(\operatorname{\mathbf{T}}\nolimits,M)^{r}\otimes(\operatorname{\mathbf{T}}\nolimits\otimes_{A^{\textrm{e}}}M)_{s}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\frown}^\scriptstyle{\widehat{\frown}}(𝐏Ae(MAN))sr\textstyle{(\operatorname{\mathbf{P}}\nolimits\otimes_{A^{\textrm{e}}}(M\otimes_{A}N))_{s-r}}(𝐓Ae(MAN))sr\textstyle{(\operatorname{\mathbf{T}}\nolimits\otimes_{A^{\textrm{e}}}(M\otimes_{A}N))_{s-r}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ϑAeidMAN\scriptstyle{\vartheta\otimes_{A^{\textrm{e}}}\operatorname{id}\nolimits_{M\otimes_{A}N}}

compatible with the differentials. Thus we have the remaining commutative squares of (2)(2). ∎

Corollary 4.5.

Let r0r\geq 0 and s1s\geq 1 be such that rs1r-s\leq-1. Then there exists a commutative diagram

Hr(A,M)H^s1(A,Nν11)\textstyle{\operatorname{H}\nolimits^{r}(A,M)\otimes\operatorname{\widehat{H}}\nolimits_{s-1}(A,{}_{1}N_{\nu^{-1}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\frown}H^sr1(A,(MAN)ν11)\textstyle{\operatorname{\widehat{H}}\nolimits_{s-r-1}(A,{}_{1}(M\otimes_{A}N)_{\nu^{-1}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H^r(A,M)H^s(A,N)\textstyle{\operatorname{\widehat{H}}\nolimits^{r}(A,M)\otimes\operatorname{\widehat{H}}\nolimits^{-s}(A,N)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}^\scriptstyle{\widehat{\smile}}H^rs(A,MAN)\textstyle{\operatorname{\widehat{H}}\nolimits^{r-s}(A,M\otimes_{A}N)}

where the vertical morphism on the right hand side is always an isomorphism, and the vertical morphism on the left hand side is an isomorphism if r0r\not=0 and an epimorphism otherwise.

Proof.

Corollary 4.2 and Proposition 4.4 imply that there exists a commutative diagram

H^r(A,M)H^s(A,N)\textstyle{\operatorname{\widehat{H}}\nolimits^{r}(A,M)\otimes\operatorname{\widehat{H}}\nolimits^{-s}(A,N)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}^\scriptstyle{\widehat{\smile}}id(^ω)\scriptstyle{\operatorname{id}\nolimits\otimes(-\widehat{\frown}\,\omega)}H^rs(A,MAN)\textstyle{\operatorname{\widehat{H}}\nolimits^{r-s}(A,M\otimes_{A}N)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}^ω\scriptstyle{-\widehat{\frown}\,\omega}H^r(A,M)H^s1(A,Nν11)\textstyle{\operatorname{\widehat{H}}\nolimits^{r}(A,M)\otimes\operatorname{\widehat{H}}\nolimits_{s-1}(A,{}_{1}N_{\nu^{-1}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}^\scriptstyle{\widehat{\frown}}H^rs(A,(MAN)ν11)\textstyle{\operatorname{\widehat{H}}\nolimits^{r-s}(A,{}_{1}(M\otimes_{A}N)_{\nu^{-1}})}Hr(A,M)H^s1(A,Nν11)\textstyle{\operatorname{H}\nolimits^{r}(A,M)\otimes\operatorname{\widehat{H}}\nolimits_{s-1}(A,{}_{1}N_{\nu^{-1}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\frown}H^sr1(A,(MAN)ν11)\textstyle{\operatorname{\widehat{H}}\nolimits_{s-r-1}(A,{}_{1}(M\otimes_{A}N)_{\nu^{-1}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

This completes the proof. ∎

Acknowledgments

The author would like to thank his PhD supervisor Professor Katsunori Sanada for giving such an interesting topic and for valuable comments and suggestions for the development of the paper. The author also would like to thank Professor Tomohiro Itagaki for helpful discussions and comments on the paper and warm encouragement.

References

  • [AM02] L. L. Avramov and A. Martsinkovsky. Absolute, relative, and Tate cohomology of modules of finite Gorenstein dimension. Proc. London Math. Soc. (3)(3), 85(2):393–440, 2002.
  • [Arm19] M. Armenta. On the cap product in Hochschild theory. arXiv:1908.02255, 2019.
  • [BGSS08] R.-O. Buchweitz, E. L. Green, N. Snashall, and Ø. Solberg. Multiplicative structures for Koszul algebras. Q. J. Math., 59(4):441–454, 2008.
  • [BJ13] P. A. Bergh and D. A. Jorgensen. Tate-Hochschild homology and cohomology of Frobenius algebras. J. Noncommut. Geom., 7(4):907–937, 2013.
  • [Bro94] K. S. Brown. Cohomology of Groups, volume 87 of Graduate Texts in Mathematics. Springer-Verlag, New York, 1994. Corrected reprint of the 1982 original.
  • [Buc86] R.-O. Buchweitz. Maximal cohen-macaulay modules and Tate-cohomology over Gorenstein rings. preprint, 1986.
  • [CJ14] L. W. Christensen and D. A. Jorgensen. Tate (co)homology via pinched complexes. Trans. Amer. Math. Soc., 366(2):667–689, 2014.
  • [ES09] C.-H. Eu and T. Schedler. Calabi-Yau Frobenius algebras. J. Algebra, 321(3):774–815, 2009.
  • [Ger63] M. Gerstenhaber. The cohomology structure of an associative ring. Ann. of Math. (2), 78:267–288, 1963.
  • [Kawse] Y. Kawada. Algebraic number theory. Kyoritsu Shuppan, Tokyo, 1957 (in Japanese).
  • [Nak57] T. Nakayama. On the complete cohomology theory of Frobenius algebras. Osaka Math. J., 9:165–187, 1957.
  • [Ngu13] V. C. Nguyen. Tate and Tate-Hochschild cohomology for finite dimensional Hopf algebras. J. Pure Appl. Algebra, 217(10):1967–1979, 2013.
  • [Ric89] J. Rickard. Derived categories and stable equivalence. J. Pure Appl. Algebra, 61(3):303–317, 1989.
  • [San92] K. Sanada. On the cohomology of Frobenius algebras. J. Pure Appl. Algebra, 80(1):65–88, 1992.
  • [SY11] A. Skowroński and K. Yamagata. Frobenius algebras. I. EMS Textbooks in Mathematics. European Mathematical Society (EMS), Zürich, 2011. Basic representation theory.
  • [Wan18] Z. Wang. Gerstenhaber algebra and Deligne’s conjecture on Tate-Hochschild cohomology. arXiv:1801.07990, 2018.
  • [Zim14] A. Zimmermann. Representation theory, volume 19 of Algebra and Applications. Springer, Cham, 2014. A homological algebra point of view.