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

Generalized parallel paths method for computing the first Hochschild cohomology groups with applications to Brauer graph algebras

Yuming Liua and Bohan Xinga,∗
Mathematics Subject Classification(2020): 16E40, 16Gxx.Keywords: Algebraic Morse theory; Brauer graph algebra; First Hochschild cohomology group; Generalized parallel paths method; Two-sided Anick resolution.aSchool of Mathematical Sciences, Laboratory of Mathematics and Complex Systems, Beijing Normal University, Beijing 100875, P. R. China. E-mail: [email protected] (Y.M. Liu); [email protected] (B.H. Xing).Corresponding author.

Abstract: We use algebraic Morse theory to generalize the parallel paths method for computing the first Hochschild cohomology groups. As an application, we describe and compare the Lie structures of the first Hochschild cohomology groups of Brauer graph algebras and their associated graded algebras.

1 Introduction

It is well-known that the Hochschild cohomology groups are important invariants of associative algebras. The computation of Hochschild cohomology groups are heavily based on a two-sided projective resolution of a given algebra. The smaller of the size of this projective resolution, the more efficient of the computation. For monomial algebras, Bardzell [Bardzell] constructed minimal two-sided projective resolutions. Based on the minimal two-sided projective resolution, Strametz [Strametz] created the parallel paths method to compute the first Hochschild cohomology group of a monomial algebra.

One aim of the present paper is to generalize Strametz’s parallel paths method on computing the first Hochschild cohomology groups from monomial algebras to general quiver algebras of the form kQ/IkQ/I where QQ is a finite quiver and II is an ideal of the path algebra kQkQ contained in kQ2kQ_{\geq 2}. Our idea is to use the two-sided Anick resolution, which is based on algebraic Morse theory, to replace Bardzell’s minimal two-sided projective resolution.

About fifteen years ago, the algebraic Morse theory was developed by Kozlov [K], by Sköldberg [ES], and by Jöllenbeck and Welker [JW], independently. Since then, this theory has been widely used in algebra; for some further references on this direction, see the introduction of the recent paper [CLZ]. In particular, Sköldberg [ES] applied this theory to construct the so-called two-sided Anick resolution from the reduced bar resolution of an non-commutative polynomial algebra. Chen, Liu and Zhou [CLZ] generalized the two-sided Anick resolution from non-commutative polynomial algebras to algebras given by quivers with relations.

For a monomial algebra kQ/IkQ/I, the ideal II has a minimal generating set ZZ given by paths in QQ, which is one of ingredients in Strametz’s construction. For arbitrary quiver algebra kQ/IkQ/I, we use the Gröbner basis of II to replace the above set ZZ. In order to generalize Strametz’s construction, we use the two-sided Anick resolution (which is for an arbitrary quiver algebra) to replace Bardzell’s minimal two-sided projective resolution (which is only for a monomial algebra). Similar to Strametz [Strametz], we also describe the Lie algebra structure on the first Hochschild cohomology group. We implement all these ideas in Section 3. It should be noted that, Artenstein, Lanzilotta and Solotar [ALS] recently studied the Hochschild cohomology of toupie algebras and the cochain complex obtained in [ALS, Section 3] to compute HH1\mathrm{HH}^{1} for toupie algebras coincides with the cochain complex in our Proposition LABEL:gen-parallel_paths.

In Section 4 we will apply our method to study the first Hochschild cohomology groups of Brauer graph algebras (or just BGAs); these algebras coincide with finite dimensional symmetric special biserial algebras. Under some mild characteristic condition (see Proposition LABEL:BGA_L-1), we can describe explicitly a set of generators of the first Hochschild cohomology group of a BGA, with a comparison in Section 5 to the first Hochschild cohomology group of the associated graded algebra. In particular, we will construct an injection ii from HH1(A)\mathrm{HH}^{1}(A) to HH1(gr(A))\mathrm{HH}^{1}(gr(A)), where AA is a BGA and gr(A)gr(A) its associated graded algebra. Actually, this map ii is always a Lie algebra monomorphism with one exception. This injection also tells us that the difference between the dimension of HH1(A)\mathrm{HH}^{1}(A) and of HH1(gr(A))\mathrm{HH}^{1}(gr(A)) is equal to the difference between the rank of Out(A)\mathrm{Out}(A)^{\circ} and of Out(gr(A))\mathrm{Out}(gr(A))^{\circ}.

After we submitted this paper on arXiv, we noticed that Rubio y Degrassi, Schroll and Solotar have recently obtained similar results in [RSS]. However, our generalized parallel paths method on computing HH1\mathrm{HH}^{1} is deduced from two-sided Anick resolutions using algebraic Morse theory, rather than directly uses the Chouhy-Solotar projective resolution which is constructed in [CS]. Moreover, we described explicitly a set of generators of HH1(A)\mathrm{HH}^{1}(A) for any Brauer graph algebra AA, and our comparison study on the first Hochschild cohomology groups between BGAs and their associated graded algebras is also new. Finally, we noticed that there would exist counter-examples of Theorem 4.2 in [RSS] in positive characteristic, see our Example LABEL:counter-example.

Outline.  In Section 2, we make some preliminaries and give some notations which we need throughout this paper; in particular, we will review the Gröbner basis theory for path algebras and the two-sided Anick resolutions for quiver algebras based on algebraic Morse theory. In Section 3 we generalize the parallel paths method for computing HH1\mathrm{HH}^{1} from monomial algebras to general quiver algebras. The main results in this section are Proposition LABEL:gen-parallel_paths and Theorem LABEL:gen-lie_bracket. In Sections 4 and 5, we give the application of the generalized parallel paths method on BGAs and their associated graded algebras; the main results are Theorems LABEL:gen-set_of_A, LABEL:grA-solvable and LABEL:inj-map. One interesting consequence (Corollary LABEL:dim(A-grA), see also Corollary LABEL:diff2) gives a simple formula for the difference between the dimensions of the first Hochschild cohomology groups of a BGA and its associated graded algebra.

2 Preliminaries

Throughout this paper we will concentrate on quiver algebras of the form kQ/IkQ/I, where kk is a field, QQ is a finite quiver, II is a two-sided ideal in the path algebra kQkQ. For each integer n0n\geq 0, we denote by QnQ_{n} the set of all paths of length nn and by QnQ_{\geq n} the set of all paths with length at least nn. We shall assume that the ideal II is contained in kQ2kQ_{\geq 2} so that kQ/IkQ/I is not necessarily finite dimensional. We denote by s(p)s(p) the source vertex of a path pp and by t(p)t(p) its terminus vertex. We will write paths from right to left, for example, p=αnαn1α1p=\alpha_{n}\alpha_{n-1}\cdots\alpha_{1} is a path with starting arrow α1\alpha_{1} and ending arrow αn\alpha_{n}. The length of a path pp will be denoted by l(p)l(p). Two paths ε,γ\varepsilon,\gamma of QQ are called parallel if s(ε)=s(γ)s(\varepsilon)=s(\gamma) and t(ε)=t(γ)t(\varepsilon)=t(\gamma). If XX and YY are sets of paths of QQ, the set X//YX//Y of parallel paths is formed by the couples (ε,γ)X×Y(\varepsilon,\gamma)\in X\times Y such that ε\varepsilon and γ\gamma are parallel paths. For instance, Q0//QnQ_{0}//Q_{n} is the set of oriented cycles of QQ of length nn. We denote by k(X//Y)k(X//Y) the kk-vector space generated by the set X//YX//Y; For a subset SS of k(X//Y)k(X//Y), we denote by S\langle S\rangle the subspace of k(X//Y)k(X//Y) generated by SS. By abuse of notation, for a subset SS of the algebra kQ/IkQ/I, we also use S\langle S\rangle to denote the ideal generated by SS.

2.1 Gröbner bases of quiver algebras

Let A=kQ/IA=kQ/I be a quiver algebra where II is generated by a set of relations. In this subsection we recall from [Green] the Gröbner basis theory for the ideal II. Let us first introduce a special kind of well-order on the basis Q0Q_{\geq 0} of the path algebra kQkQ. By [Green, Section 2.2.2], a well-order >> on Q0Q_{\geq 0} is called admissible if it satisfies the following conditions where p,q,r,sQ0p,q,r,s\in Q_{\geq 0}:

  • if p<qp<q then pr<qrpr<qr if both pr0pr\neq 0 and qr0qr\neq 0.

  • if p<qp<q then sp<sqsp<sq if both sp0sp\neq 0 and sq0sq\neq 0.

  • if p=qrp=qr, then pqp\geq q and prp\geq r.

Given a quiver QQ as above, there are “natural” admissible orders on Q0Q_{\geq 0}. Here is one example:

Example 2.1.

The (left) length-lexicographic order on Q0Q_{\geq 0}:

Order the vertices Q0={v1,,vn}Q_{0}=\{v_{1},\cdots,v_{n}\} and arrows Q1={a1,,am}Q_{1}=\{a_{1},\cdots,a_{m}\} arbitrarily and set the vertices smaller than the arrows. Thus

v1<<vn<a1<<am.v_{1}<\cdots<v_{n}<a_{1}<\cdots<a_{m}.

If pp and qq are paths of length at least 11, set p<qp<q if l(p)<l(q)l(p)<l(q) or if p=b1brp=b_{1}\cdots b_{r} and q=b1brq=b_{1}^{\prime}\cdots b_{r}^{\prime} with b1,,br,b1,,brQ1b_{1},\cdots,b_{r},b_{1}^{\prime},\cdots,b_{r}^{\prime}\in Q_{1} and for some 1ir1\leq i\leq r, bj=bjb_{j}=b_{j}^{\prime} for j<ij<i and bi<bib_{i}<b_{i}^{\prime}.

We now fix an admissible well-order >> on Q0Q_{\geq 0}. For any akQa\in kQ, we have a=pQ0,λpkλppa=\sum_{p\in Q_{\geq 0},\;\lambda_{p}\in k}\lambda_{p}p and write Supp(a)={pλp0}\mathrm{Supp}(a)=\{p\mid\lambda_{p}\neq 0\}. We call Tip(a)=p\mathrm{Tip}(a)=p, if pSupp(a)p\in\mathrm{Supp}(a) and ppp^{\prime}\leq p for all pSupp(a)p^{\prime}\in\mathrm{Supp}(a). Then we denote the tip of a set WkQW\subseteq kQ by Tip(W)={Tip(w)|wW}\mathrm{Tip}(W)=\{\mathrm{Tip}(w)|w\in W\} and write NonTip(W):=Q0\Tip(W)\mathrm{NonTip}(W):=Q_{\geq 0}\backslash\mathrm{Tip}(W). We also denote the coefficient of the tip of aa by CTip(a)\mathrm{CTip}(a). In particular, we will use Tip(I)\mathrm{Tip}(I) and NonTip(I)\mathrm{NonTip}(I) for the ideal II of kQkQ. By [Green], there is a decomposition of vector spaces

kQ=ISpank(NonTip(I)).kQ=I\oplus\mathrm{Span}_{k}(\mathrm{NonTip}(I)).

So NonTip(I)\mathrm{NonTip}(I) (modulo II) gives a “monomial” kk-basis of the quotient algebra A=kQ/IA=kQ/I.

Definition 2.2.

([Green,  Definition 2.4])(\cite[cite]{[\@@bibref{}{Green}{}{}, ~{}Definition~{}2.4]}) With the notations as above, a subset 𝒢I\mathcal{G}\subseteq I is a Gröbner basis for the ideal II with respect to the order >> if

Tip(I)=Tip(𝒢),\langle\mathrm{Tip}(I)\rangle=\langle\mathrm{Tip}(\mathcal{G})\rangle,

that is, Tip(I)\mathrm{Tip}(I) and Tip(𝒢)\mathrm{Tip}(\mathcal{G}) generate the same ideal in kQkQ.

Actually, in this case I=𝒢I=\langle\mathcal{G}\rangle. By the discussion in [Green], we have a complete method to judge whether a set of generators of an ideal II in kQkQ is a Gröbner basis, which is called the Termination Theorem. The idea is to check whether some special elements of the ideal II are divisible by this basis, instead of to check all the elements in II.

Definition 2.3.

([Green,  Definition 2.7])(\cite[cite]{[\@@bibref{}{Green}{}{}, ~{}Definition~{}2.7]}) Let kQkQ be a path algebra, >> an admissible order on Q0Q_{\geq 0} and f,gkQf,g\in kQ. Suppose b,cQ0b,c\in Q_{\geq 0}, such that

  • Tip(f)c=bTip(g)\mathrm{Tip}(f)c=b\mathrm{Tip}(g),

  • Tip(f)b\mathrm{Tip}(f)\nmid b, Tip(g)c\mathrm{Tip}(g)\nmid c.

Then the overlap relation of ff and gg by b,cb,c is

o(f,g,b,c)=(CTip(f))1fc(CTip(g))1bg.o(f,g,b,c)=(\mathrm{CTip}(f))^{-1}\cdot fc-(\mathrm{CTip}(g))^{-1}\cdot bg.

Note that Tip(o(f,g,b,c))<Tip(f)c=bTip(g)\mathrm{Tip}(o(f,g,b,c))<\mathrm{Tip}(f)c=b\mathrm{Tip}(g).

Theorem 2.4.

([Green,  Theorem 2.3])(\cite[cite]{[\@@bibref{}{Green}{}{}, ~{}Theorem~{}2.3]}) Let kQkQ be a path algebra, >> an admissible order on Q0Q_{\geq 0}, 𝒢\mathcal{G} a set of elements of kQkQ. Suppose for every overlap relation, we have

o(g1,g2,p,q)𝒢0,o(g_{1},g_{2},p,q)\Rightarrow_{\mathcal{G}}0,

which means that o(g1,g2,p,q)o(g_{1},g_{2},p,q) can be divided by Tip(𝒢)\mathrm{Tip}(\mathcal{G}), with g1,g2𝒢g_{1},g_{2}\in\mathcal{G} and p,qQ0p,q\in Q_{\geq 0}. Then 𝒢\mathcal{G} is a Gröbner basis of 𝒢\langle\mathcal{G}\rangle, the ideal generated by 𝒢\mathcal{G}.

Definition 2.5.

([Green,  Definition 2.8])(\cite[cite]{[\@@bibref{}{Green}{}{}, ~{}Definition~{}2.8]}) A Gröbner basis 𝒢\mathcal{G} for the ideal II is reduced if the following three conditions are satisfied:

  • 𝒢\mathcal{G} is tip-reduced: for g,h𝒢g,h\in\mathcal{G} with ghg\neq h, Tip(g)Tip(h)\mathrm{Tip}(g)\nmid\mathrm{Tip}(h);

  • 𝒢\mathcal{G} is monic: for every element g𝒢g\in\mathcal{G}, CTip(g)=1\mathrm{CTip}(g)=1;

  • For any g𝒢g\in\mathcal{G}, gTip(g)Spank(NonTip(I))g-\mathrm{Tip}(g)\in\mathrm{Span}_{k}(\mathrm{NonTip}(I)).

It is easy to see that under a given admissible order, II has a unique reduced Gröbner basis 𝒢\mathcal{G}, and in this case Tip(𝒢)\mathrm{Tip}(\mathcal{G}) is a minimal generator set of Tip(I)\langle\mathrm{Tip}(I)\rangle; moreover, bQ0b\in Q_{\geq 0} lies in NonTip(I)\mathrm{NonTip}(I) if and only if bb is not divided by any element of Tip(𝒢)\mathrm{Tip}(\mathcal{G}). In the following, we always assume that 𝒢\mathcal{G} is a reduced Gröbner basis of II.

Note that when 𝒢\mathcal{G} is reduced, there is a one-to-one correspondence between 𝒢\mathcal{G} and Tip(𝒢)\mathrm{Tip}(\mathcal{G}): for g𝒢g\in\mathcal{G}, Tip(g)Tip(𝒢)\mathrm{Tip}(g)\in\mathrm{Tip}(\mathcal{G}); conversely, for wTip(𝒢)w\in\mathrm{Tip}(\mathcal{G}), there is a unique g𝒢g\in\mathcal{G} such that w=Tip(g)w=\mathrm{Tip}(g). We shall denote the correspondence from 𝒢\mathcal{G} to Tip(𝒢)\mathrm{Tip}(\mathcal{G}) by Tip\mathrm{Tip} and its inverse by Tip1\mathrm{Tip^{-1}}.

2.2 The reduced bar resolution of quiver algebras

Now let E:eQ0keE:\simeq\oplus_{e\in Q_{0}}ke be the separable subalgebra of AA generated by the classes modulo II of the vertices of QQ, such that A=EA+A=E\oplus A_{+} as EeE^{e}-modules, where Ee=EkEE^{e}=E\otimes_{k}E and A+:=Spank{NonTip(I)\Q0}A_{+}:=\mathrm{Span}_{k}\{\mathrm{NonTip}(I)\backslash Q_{0}\}. Actually, there is an EeE^{e}-projection from AA to A+A_{+}, denoted by pAp_{A}. The reduced bar resolution of the quiver algebra AA can be written by the form of the following theorem in the sense of Cibils [C].

Theorem 2.6.

For the algebra A=kQ/IA=kQ/I, the reduced bar resolution B(A)B(A) is a two-sided projective resolution of AA with B0(A)=AEAB_{0}(A)=A\otimes_{E}A, Bn(A)=AE(A+)EnEAB_{n}(A)=A\otimes_{E}(A_{+})^{\otimes_{E}n}\otimes_{E}A and the differential d=(dn)d=(d_{n}) is

dn([a1||an])=a1[a2||an]+i=1n1(1)i[a1||aiai+1||an]+(1)n[a1||an1]and_{n}([a_{1}|\cdots|a_{n}])=a_{1}[a_{2}|\cdots|a_{n}]+\sum_{i=1}^{n-1}(-1)^{i}[a_{1}|\cdots|a_{i}a_{i+1}|\cdots|a_{n}]+(-1)^{n}[a_{1}|\cdots|a_{n-1}]a_{n}

with [a1||an]:=1a1an1[a_{1}|\cdots|a_{n}]:=1\otimes a_{1}\otimes\cdots\otimes a_{n}\otimes 1. By convention B1(A)=AB_{-1}(A)=A, and d0:AEAAd_{0}:A\otimes_{E}A\longrightarrow A is given by the multiplication μA\mu_{A} in AA.

Remark 2.7.

By the definition of A+A_{+}, Bn(A)B_{n}(A) can be decomposed as

Bn(A)=A[w1||wn]A=Ae[w1||wn],B_{n}(A)=\bigoplus A[w_{1}|\cdots|w_{n}]A=\bigoplus A^{e}[w_{1}|\cdots|w_{n}],

where Ae=AkAopA^{e}=A\otimes_{k}A^{op} is the enveloping algebra of AA and the direct sum is taken over all the signs [w1||wn][w_{1}|\cdots|w_{n}] such that all wiNonTip(I)\Q0w_{i}\in\mathrm{NonTip}(I)\backslash Q_{0} and w1wnw_{1}\cdots w_{n} is a path in QQ.

Since the reduced bar resolution is a two-sided projective resolution of AA, we can use it to compute the Hochschild cohomology groups HHn(A)\mathrm{HH}^{n}(A) of AA. More concretely, applying the functor HomAe(,A)\mathrm{Hom}_{{A}^{e}}(-,{A}) to B(A)B(A) we get a cochain complex (C(A),δ)(C^{*}(A),\delta^{*}), where C0(A)HomEe(E,A)AE={aAsa=as for all sE}C^{0}(A)\cong\mathrm{Hom}_{E^{e}}(E,A)\cong A^{E}=\{a\in A\mid sa=as\text{ for all }s\in E\}, Cn(A)=HomAe(Bn(A),A)HomEe((A+)En,A)C^{n}(A)=\mathrm{Hom}_{A^{e}}(B_{n}(A),A)\cong\mathrm{Hom}_{E^{e}}((A_{+})^{\otimes_{E}n},A) for n1n\geq 1 (cf. Lemma 3.3), (δ0a)(x)=axxa(\delta^{0}a)(x)=ax-xa for aAEa\in A^{E} and xA+x\in A_{+}, and

(δnf)(x1xn+1)=x1f(x2xn+1)(\delta^{n}f)(x_{1}\otimes\cdots\otimes x_{n+1})=x_{1}f(x_{2}\otimes\cdots\otimes x_{n+1})
+i=1n(1)if(x1xixi+1xn+1)+(1)n+1f(x1xn)xn+1.+\sum_{i=1}^{n}(-1)^{i}f(x_{1}\otimes\cdots\otimes x_{i}x_{i+1}\otimes\cdots\otimes x_{n+1})+(-1)^{n+1}f(x_{1}\otimes\cdots\otimes x_{n})x_{n+1}.

Then we have

HHn(A)=Kerδn/Imδn1.\mathrm{HH}^{n}(A)=\mathrm{Ker}\delta^{n}/\mathrm{Im}\delta^{n-1}.

In particular, we have HH1(A)=Kerδ1/Imδ0\mathrm{HH}^{1}(A)=\mathrm{Ker}\delta^{1}/\mathrm{Im}\delta^{0} as kk-spaces, where Kerδ1\mathrm{Ker}\delta^{1} is the set of EeE^{e}-derivations of A+A_{+} into AA and the elements in Imδ0\mathrm{Im}\delta^{0} are inner EeE^{e}-derivations of A+A_{+} into AA. Note that we can identify DerEe(A+,A)\mathrm{Der}_{E^{e}}(A_{+},A) with DerEe(A,A)\mathrm{Der}_{E^{e}}(A,A) and Kerδ1=DerEe(A+,A)\mathrm{Ker}\delta^{1}=\mathrm{Der}_{E^{e}}(A_{+},A) has a Lie algebra structure under the Lie bracket

[f,g]HH:=fpAggpAf[f,g]_{HH}:=f\circ p_{A}\circ g-g\circ p_{A}\circ f

for f,gDerEe(A+,A)f,g\in\mathrm{Der}_{E^{e}}(A_{+},A), where pAp_{A} denotes the EeE^{e}-projection from AA to A+A_{+}. Moreover, Imδ0\mathrm{Im}\delta^{0} is a Lie ideal of Kerδ1\mathrm{Ker}\delta^{1}, so that HH1(A)\mathrm{HH}^{1}(A) is a Lie algebra. This structure was first defined by Gerstenhaber [Ger] using the standard bar resolution of AA.

In next two subsections we will explain how to use the algebraic Morse theory to shrink the above reduced bar resolution of AA to a “smaller” one, such that the homology of the two complexes coincides.

2.3 Algebraic Morse theory

The most general version of algebraic Morse theory was presented in Chen, Liu and Zhou [CLZ]. For our purpose, we will adopt to a Morse matching condition defined in [CLZ, Proposition 3.2].

Let RR be an associative ring and C=(Cn,n)nC_{*}=(C_{n},\partial_{n})_{n\in\mathbb{Z}} be a chain complex of left RR-modules. We assume that each RR-module CnC_{n} has a decomposition CniInCn,iC_{n}\simeq\oplus_{i\in I_{n}}C_{n,i} of RR-modules, so we can regard the differentials n\partial_{n} as a matrix n=(n,ji)\partial_{n}=(\partial_{n,ji}) with iIni\in I_{n} and jIn1j\in I_{n-1} and where n,ji:Cn,iCn1,j\partial_{n,ji}:C_{n,i}\rightarrow C_{n-1,j} is a homomorphism of RR-modules.

Given the complex CC_{*} as above, we construct a weighted quiver G(C):=(V,E)G(C_{*}):=(V,E). The set VV of vertices of G(C)G(C_{*}) consists of the pairs (n,i)(n,i) with n,iInn\in\mathbb{Z},i\in I_{n} and the set EE of weighted arrows is given by the rule: if the map n,ji\partial_{n,ji} does not vanish, draw an arrow in E from (n,i)(n,i) to (n1,j)(n-1,j) and denote the weight of this arrow by the map n,ji\partial_{n,ji}.

A full subquiver \mathcal{M} of the weighted quiver G(C)G(C_{*}) is called a partial matching if it satisfies the following two conditions:

  • (Matching)(Matching) Each vertex in VV belongs to at most one arrow of \mathcal{M}.

  • (Invertibility)(Invertibility) Each arrow in \mathcal{M} has its weight invertible as a RR-homomorphism.

With respect to a partial matching \mathcal{M}, we can define a new weighted quiver G(C)=(V,E)G_{\mathcal{M}}(C_{*})=(V,E_{\mathcal{M}}), where EE_{\mathcal{M}} is given by

  • Keep everything for all arrows which are not in \mathcal{M} and call them thick arrows.

  • For an arrow in \mathcal{M}, replace it by a new dotted arrow in the reverse direction and the weight of this new arrow is the negative inverse of the weight of original arrow.

A path in G(C)G_{\mathcal{M}}(C_{*}) is called a zigzag path if dotted arrows and thick arrows appear alternately.

Next, for convenience, we will introduce from Jöllenbeck and Welker [JW] the notations related to the weighted quiver G(C)=(V,E)G(C_{*})=(V,E) with a partial matching \mathcal{M} on it.

Definition 2.8.
  1. (1)

    A vertex (n,i)V(n,i)\in V is critical with respect to \mathcal{M} if (n,i)(n,i) does not lie in any arrow in \mathcal{M}. Let VnV_{n} denote all the vertices with the first number equal to nn, we write

    Vn:={(n,i)Vn|(n,i)iscritical}V_{n}^{\mathcal{M}}:=\{(n,i)\in V_{n}\;|\;(n,i)\;is\;critical\}

    for the set of all critical vertices of homological degree nn.

  2. (2)

    Write (m,j)(n,i)(m,j)\leq(n,i) if there exists an arrow from (n,i)(n,i) to (m,j)(m,j) in G(C)G(C_{*}).

  3. (3)

    Denote by P((n,i),(m,j))P((n,i),(m,j)) the set of all zigzag paths from (n,i)(n,i) to (m,j)(m,j) in G(C)G_{\mathcal{M}}(C_{*}).

  4. (4)

    The weight w(p)w(p) of a path

    p=((n1,i1)(n2,i2)(nr,ir))P((n1,i1),(nr,ir))p=((n_{1},i_{1})\rightarrow(n_{2},i_{2})\rightarrow\cdots\rightarrow(n_{r},i_{r}))\in P((n_{1},i_{1}),(n_{r},i_{r}))

    in G(C)G_{\mathcal{M}}(C_{*}) is given by

    w(p):=w((nr1,ir1)(nr,ir))w((n1,i1)(n2,i2)),w(p):=w((n_{r-1},i_{r-1})\rightarrow(n_{r},i_{r}))\circ\cdots\circ w((n_{1},i_{1})\rightarrow(n_{2},i_{2})),
    w((n,i)(m,j)):={m,ij1,(n,i)(m,j),n,ji,(m,j)(n,i).w((n,i)\rightarrow(m,j)):=\left\{\begin{array}[]{*{3}{lll}}-\partial_{m,ij}^{-1}&,&(n,i)\leq(m,j),\\ \partial_{n,ji}&,&(m,j)\leq(n,i).\end{array}\right.

    Then we write Γ((n,i),(m,j))=pP((n,i),(m,j))w(p)\Gamma((n,i),(m,j))=\sum_{p\in P((n,i),(m,j))}w(p) for the sum of weights of all zigzag paths from (n,i)(n,i) to (m,j)(m,j).

Following [CLZ, Proposition 3.2], we call a partial matching \mathcal{M} as above a Morse matching if any zigzag path starting from (n,i)(n,i) is of finite length for each vertex (n,i)(n,i) in G(C)G_{\mathcal{M}}(C_{*}).

Now we can define a new complex CC_{*}^{\mathcal{M}}, which we call the Morse complex of CC_{*} with respect to \mathcal{M}. The complex C=(Cn,n)nC_{*}^{\mathcal{M}}=(C_{n}^{\mathcal{M}},\partial_{n}^{\mathcal{M}})_{n\in\mathbb{Z}} is defined by

Cn:=(n,i)VnCn,i,C_{n}^{\mathcal{M}}:=\oplus_{(n,i)\in V_{n}^{\mathcal{M}}}C_{n,i},
n:{CnCn1xCn,i(n1,j)Vn1Γ((n,i),(n1,j))(x).\partial_{n}^{\mathcal{M}}:\left\{\begin{array}[]{*{3}{lll}}C_{n}^{\mathcal{M}}&\rightarrow&C_{n-1}^{\mathcal{M}}\\ x\in C_{n,i}&\mapsto&\sum_{(n-1,j)\in V_{n-1}^{\mathcal{M}}}\Gamma((n,i),(n-1,j))(x).\end{array}\right.

The main theorem of algebraic Morse theory can be stated as follows.

Theorem 2.9.

CC_{*}^{\mathcal{M}} is a complex of left RR-modules which is homotopy equivalent to the original complex CC_{*}. Moreover, the maps defined below are chain homotopies between CC_{*} and CC_{*}^{\mathcal{M}}:

f:{CnCnxCn,i(n,j)VnΓ((n,i),(n,j))(x),f:\left\{\begin{array}[]{*{3}{lll}}C_{n}&\rightarrow&C_{n}^{\mathcal{M}}\\ x\in C_{n,i}&\mapsto&\sum_{(n,j)\in V_{n}^{\mathcal{M}}}\Gamma((n,i),(n,j))(x),\end{array}\right.
g:{CnCnxCn,i(n,j)VnΓ((n,i),(n,j))(x).g:\left\{\begin{array}[]{*{3}{lll}}C_{n}^{\mathcal{M}}&\rightarrow&C_{n}\\ x\in C_{n,i}&\mapsto&\sum_{(n,j)\in V_{n}}\Gamma((n,i),(n,j))(x).\end{array}\right.

2.4 Two-sided Anick resolution

Starting from the reduced bar resolution of an one-vertex quiver algebra AA which is viewed as a chain complex of projective AeA^{e}-modules, Sköldberg [ES] constructed a “smaller” AeA^{e}-projective resolution of AA using algebraic Morse theory, which is called the two-sided Anick resolution of AA. It was pointed out in [CLZ] that Sköldberg’s construction generalizes to general quiver algebras.

Let A=kQ/IA=kQ/I be a quiver algebra, let 𝒢\mathcal{G} be a reduced Gröbner basis of the ideal II, and denote W:=Tip(𝒢)W:=\mathrm{Tip}(\mathcal{G}). Denote by B(A)=(B(A),d)B(A)=(B_{*}(A),d_{*}) the reduced bar resolution of AA (cf. Section 2.2). Similar as in [CLZ], we define a new quiver QW=(V,E)Q_{W}=(V,E) with respect to WW, which is called the Ufnarovskiĭ graph (or just Uf-graph).

Definition 2.10.

A Uf-graph QW=(V,E)Q_{W}=(V,E) with respect to WW of the algebra A=kQ/IA=kQ/I is given by

V:=Q0Q1{uQ0|u is a proper right factor of some vW}V:=Q_{0}\cup Q_{1}\cup\{u\in Q_{\geq 0}\ |\ \text{$u$ is a proper right factor of some $v\in W$}\}
E:={ex|eQ0,x=exQ1}E:=\{e\rightarrow x\ |\ e\in Q_{0},\;x=ex\in Q_{1}\}~{}~{}\cup
{uv|uvTip(𝒢), but wTip(𝒢)foruv=wp,l(p)1}\{u\rightarrow v\ |\ uv\in\langle\mathrm{Tip}(\mathcal{G})\rangle,\text{ but }w\notin\langle\mathrm{Tip}(\mathcal{G})\rangle\;for\;uv=wp,\;l(p)\geq 1\}

Using Uf-graph QWQ_{W} one can define (for each i1i\geq-1) the ii-chains, which form a subset of generators of Bi(A)=Ae[w1||wi]B_{i}(A)=\bigoplus A^{e}[w_{1}|\cdots|w_{i}] for i0i\geq 0, with w1,,wiNonTip(I)\Q0w_{1},\cdots,w_{i}\in\mathrm{NonTip}(I)\backslash Q_{0} and w1wiQ0w_{1}\cdots w_{i}\in Q_{\geq 0}.

Definition 2.11.
  • The set W(i)W^{(i)} of ii-chains consists of all sequences [w1||wi+1][w_{1}|\cdots|w_{i+1}] with each wkNonTip(I)\Q0w_{k}\in\mathrm{NonTip}(I)\backslash Q_{0}, such that

    ew1w2wi+1e\rightarrow w_{1}\rightarrow w_{2}\rightarrow\cdots\rightarrow w_{i+1}

    is a path in QWQ_{W}. And define W(1):=Q0W^{(-1)}:=Q_{0}.

  • For for all pQ0p\in Q_{\geq 0}, define

    Vp,i(n)={[w1||wn]|p=w1wn,[w1||wi+1]W(i),[w1||wi+2]W(i+1)}V_{p,i}^{(n)}=\{[w_{1}|\cdots|w_{n}]\ |\ p=w_{1}\cdots w_{n},\;[w_{1}|\cdots|w_{i+1}]\in W^{(i)},\;[w_{1}|\cdots|w_{i+2}]\notin W^{(i+1)}\}

By using the definition above, we can define a partial matching \mathcal{M} to be the set of arrows of the following form in the weighted quiver G(B)G(B_{*}), where B=B(A)B_{*}=B(A) is the reduced bar resolution of the algebra A=kQ/IA=kQ/I:

[w1||wi+1|wi+2|wi+2′′|wi+3||wn](1)i+2[w1||wi+2||wn][w_{1}|\cdots|w_{i+1}|w_{i+2}^{\prime}|w_{i+2}^{\prime\prime}|w_{i+3}|\cdots|w_{n}]\stackrel{{\scriptstyle(-1)^{i+2}}}{{\longrightarrow}}[w_{1}|\cdots|w_{i+2}|\cdots|w_{n}]

where

w=w1wn=w1wi+2wi+2′′wn,wi+2=wi+2wi+2′′,w=w_{1}\cdots w_{n}=w_{1}\cdots w_{i+2}^{\prime}w_{i+2}^{\prime\prime}\cdots w_{n},~{}~{}w_{i+2}=w_{i+2}^{\prime}w_{i+2}^{\prime\prime},
[w1||wi+1|wi+2|wi+2′′|wi+3||wn]Vw,i+1(n+1),[w1||wi+2||wn]Vw,i(n).[w_{1}|\cdots|w_{i+1}|w_{i+2}^{\prime}|w_{i+2}^{\prime\prime}|w_{i+3}|\cdots|w_{n}]\in V_{w,i+1}^{(n+1)},~{}~{}[w_{1}|\cdots|w_{i+2}|\cdots|w_{n}]\in V_{w,i}^{(n)}.
Theorem 2.12.

([CLZ,  Theorem 4.3])(\cite[cite]{[\@@bibref{}{CLZ}{}{}, ~{}Theorem~{}4.3]}) The partial matching \mathcal{M} is a Morse matching of G(B)G(B_{*}) with the set of critical vertices in nn-th component is W(n1)W^{(n-1)}.

Therefore, by Theorem 2.12 and Theorem 2.9, (B(A),d)(B^{\mathcal{M}}(A),d^{\mathcal{M}}) is also a AeA^{e}-projective resolution of AA, but the projective AeA^{e}-modules in B(A)B^{\mathcal{M}}(A) are usually much smaller than B(A)B(A). The resolution (B(A),d)(B^{\mathcal{M}}(A),d^{\mathcal{M}}) is called the two-sided Anick resolution of AA in [ES] and [CLZ].

Note that for n0n\geq 0, the nn-th component of the two-sided Anick resolution B(A)B^{\mathcal{M}}(A) is AEkW(n1)EAA\otimes_{E}kW^{(n-1)}\otimes_{E}A. In particular, we have the following identifications:

W(1)=Q0,W(0)={[w1]|w1Q1}Q1,W^{(-1)}=Q_{0},~{}~{}W^{(0)}=\{[w_{1}]\ |\ w_{1}\in Q_{1}\}\cong Q_{1},
W(1)={[w1|w2]|w1Q1,w1w2Tip(𝒢)}Tip(𝒢)=W.W^{(1)}=\{[w_{1}|w_{2}]\ |\ \;w_{1}\in Q_{1},\;w_{1}w_{2}\in\mathrm{Tip}(\mathcal{G})\}\cong\mathrm{Tip}(\mathcal{G})=W.

3 Generalized parallel paths method for computing the first Hochschild cohomology group

3.1 Parallel paths method of Strametz

Let A=kQ/I{A}=kQ/I be a quiver algebra and EkQ0E\simeq kQ_{0} be its separable subalgebra as in Section 2.2. Remind that we assume IkQ2I\subseteq kQ_{\geq 2}. The following notations (most of them are taken from [Strametz]) are useful.

Definition 3.1.
  • Let ε\varepsilon be a path in QQ and (α,γ)Q1//Q(\alpha,\gamma)\in Q_{1}//Q. Denote by ε(α,γ)\varepsilon^{(\alpha,\gamma)} the sum of all nonzero paths obtained by replacing one appearance of the arrow α\alpha in ε\varepsilon by path γ\gamma. If the path ε\varepsilon does not contain the arrow α\alpha, we set ε(α,γ)=0\varepsilon^{(\alpha,\gamma)}=0.

  • If we fix a Gröbner basis 𝒢\mathcal{G} of II in kQkQ, then there is a kk-linear basis \mathcal{B} of the algebra A{A} with respect to 𝒢\mathcal{G}. Indeed \mathcal{B} is given by NonTip(𝒢)\mathrm{NonTip}(\mathcal{G}) modulo II and there is a bijection between the elements of \mathcal{B} and the elements of NonTip(𝒢)\mathrm{NonTip}(\mathcal{G}) (cf. Section 2.1). By this reason, we often identify \mathcal{B} with NonTip(𝒢)\mathrm{NonTip}(\mathcal{G}).

  • Let the canonical projection be written as π:kQA\pi:kQ\rightarrow{A}. For a path pQp\in Q, π(p)\pi(p) can be uniquely written as a linear combination of the elements in the basis \mathcal{B}.

  • If XX is a set of paths of QQ and ee a vertex of QQ, the set XeXe is formed by the paths of XX with source vertex ee. In the same way eXeX denotes the set of all paths of XX with terminus vertex ee.

We now give a brief review about Strametz’s method in [Strametz] for computing the first Hochschild cohomology groups of monomial algebras. Recall that A=kQ/I{A}=kQ/I is a monomial algebra means that the ideal II is generated by a set ZZ of paths in QQ. We shall assume that ZZ is minimal, so that ZZ is a reduced Gröbner basis of II with Z=Tip(Z)Z=\mathrm{Tip}(Z). Note that NonTip(Z)\mathrm{NonTip}(Z) modulo II gives a kk-basis of AA, which we denote by \mathcal{B}.

Proposition 3.2.

([Strametz,  Proposition 2.6])(\cite[cite]{[\@@bibref{}{Strametz}{}{}, ~{}Proposition~{}2.6]}) Let AA be a finite dimensional monomial algebra. Then the beginning of the cochain complex of the minimal AeA^{e}-projective resolution of AA can be described by:

0{0}k(Q0//){{k(Q_{0}//\mathcal{B})}}k(Q1//){{k(Q_{1}//\mathcal{B})}}k(Z//){{k(Z//\mathcal{B})}}{\cdots}ψ0\scriptstyle{\psi_{0}}ψ1\scriptstyle{\psi_{1}}

where the differentials are given by

ψ0:k(Q0//)k(Q1//),(e,γ)αQ1e(α,π(αγ))βeQ1(β,π(γβ));ψ1:k(Q1//)k(Z//),(α,γ)pZ(p,π(p(α,γ))).\begin{array}[]{*{5}{lllll}}\psi_{0}&:&k(Q_{0}//\mathcal{B})&\rightarrow&k(Q_{1}//\mathcal{B}),\\ &&(e,\gamma)&\mapsto&\sum_{\alpha\in Q_{1}e}(\alpha,\pi(\alpha\gamma))-\sum_{\beta\in eQ_{1}}(\beta,\pi(\gamma\beta));\\ \\ \psi_{1}&:&k(Q_{1}//\mathcal{B})&\rightarrow&k(Z//\mathcal{B}),\\ &&(\alpha,\gamma)&\mapsto&\sum_{p\in Z}(p,\pi(p^{(\alpha,\gamma)})).\end{array}

In particular, we have HH0(A)Kerψ0\mathrm{HH}^{0}({A})\cong\mathrm{Ker}\psi_{0}, HH1(A)Kerψ1/Imψ0\mathrm{HH}^{1}({A})\cong\mathrm{Ker}\psi_{1}/\mathrm{Im}\psi_{0}.

The proof of Proposition 3.2 uses the minimal AeA^{e}-projective resolution of the monomial algebra AA given by Bardzell [Bardzell] and the following two general lemmas.

Lemma 3.3.

([Strametz,  Lemma 2.2])(\cite[cite]{[\@@bibref{}{Strametz}{}{}, ~{}Lemma~{}2.2]}) Let A=kQ/I{A}=kQ/I be a quiver algebra, EkQ0E\simeq kQ_{0} be the separable subalgebra of AA, MM be an EE-bimodule and TT be a A{A}-bimodule. Then the vector space HomAe(AEMEA,T)\mathrm{Hom}_{{A}^{e}}({A}\otimes_{E}M\otimes_{E}{A},T) is isomorphic to HomEe(M,T)\mathrm{Hom}_{E^{e}}(M,T).

Proof.

It is easy to check that the kk-linear maps given by

f(mf(1m1))f\mapsto{(m\mapsto f(1\otimes m\otimes 1))}

and

g(1m1g(m))g\mapsto{(1\otimes m\otimes 1\mapsto g(m))}

are well-defined and inverse to each other. ∎

Lemma 3.4.

([Strametz,  Lemma 2.3])(\cite[cite]{[\@@bibref{}{Strametz}{}{}, ~{}Lemma~{}2.3]}) Let A=kQ/I{A}=kQ/I be a quiver algebra and EkQ0E\simeq kQ_{0} be its separable subalgebra. Let XX and YY be the sets of paths of QQ and let kXkX and kYkY be the corresponding EE-bimodules. If XX is a finite set, then the vector spaces k(X//Y)k(X//Y) and HomEe(kX,kY)\mathrm{Hom}_{E^{e}}(kX,kY) are isomorphic.

Proof.

It is easy to check that the kk-linear maps given by

(x,y)(xy,x0, for xx)(x,y)\mapsto{(x\mapsto y,\;x^{\prime}\mapsto 0,\text{ for }x^{\prime}\neq x)}

and

fxXiλi(x,pi)f\mapsto\sum_{x\in X}\sum_{i}\lambda_{i}(x,p_{i})

with f(x)=iλipif(x)=\sum_{i}\lambda_{i}p_{i}, λik\lambda_{i}\in k, piYp_{i}\in Y, are well-defined and inverse to each other. ∎

Remark 3.5.

The lemma above is the intrinsic reason why the parallel paths method can not always be used for the quiver algebras with infinite dimension, since we need to restrict XX to be finite to make the maps above well-defined. In particular, for X=Tip𝒢X=\mathrm{Tip}\mathcal{G}, in order to use above lemma, we need it to be a finite set. When kQ/IkQ/I is finite dimensional, Tip𝒢\mathrm{Tip}\mathcal{G} is always a finite set by [GL,  Proposition 2.10]. When kQ/IkQ/I is infinite dimensional but II owns a finite Gröbner basis, we can still use above lemma for X=Tip𝒢X=\mathrm{Tip}\mathcal{G}. For example, we can use the above lemma to X=Tip𝒢X=\mathrm{Tip}\mathcal{G} for the algebra kx,y/x2k\langle x,y\rangle/\langle x^{2}\rangle, but not for the algebra kx,y/xyx,xyyx,xyyyx,k\langle x,y\rangle/\langle xyx,xyyx,xyyyx,\cdots\rangle.

3.2 Generalized parallel paths method

In this subsection, we will extend parallel paths method for computing the first Hochschild cohomology groups from monomial algebras to general quiver algebras, which we call the generalized parallel paths method.

Let A=kQ/I{A}=kQ/I be a quiver algebra such that II has a finite reduced Gröbner basis 𝒢\mathcal{G}. The following lemma can be seen as a generalization of the beginning of the two-sided minimal projective resolution of a monomial algebra given by Bardzell [Bardzell]. Our proof is a careful analysis of the beginning of the two-sided Anick resolution (cf. Section 2.4) of AA.

Lemma 3.6.

The beginning of the two-sided Anick resolution (B(A),d)(B^{\mathcal{M}}({A}),d^{\mathcal{M}}) of A{A} can be described by (for simplicity we just denote dd^{\mathcal{M}} by dd):

{\cdots}AETip(𝒢)EA{{{A}\otimes_{E}\mathrm{Tip}(\mathcal{G})\otimes_{E}{A}}}AEQ1EA{{{A}\otimes_{E}Q_{1}\otimes_{E}{A}}}AEQ0EA{{{A}\otimes_{E}Q_{0}\otimes_{E}{A}}}A{{A}}0,{0,}d1\scriptstyle{d_{1}}d0\scriptstyle{d_{0}}d2\scriptstyle{d_{2}}

where the differentials can be described as follows:

  • for all 1ei1AEQ0EA1\otimes e_{i}\otimes 1\in{A}\otimes_{E}Q_{0}\otimes_{E}{A}, d0(1ei1)=eid_{0}(1\otimes e_{i}\otimes 1)=e_{i};

  • for all 1α1AEQ1EA1\otimes\alpha\otimes 1\in{A}\otimes_{E}Q_{1}\otimes_{E}{A}, d1(1α1)=αs(α)11t(α)αd_{1}(1\otimes\alpha\otimes 1)=\alpha\otimes s(\alpha)\otimes 1-1\otimes t(\alpha)\otimes\alpha;

  • for all 1w1AETip(𝒢)EA1\otimes w\otimes 1\in{A}\otimes_{E}\mathrm{Tip}(\mathcal{G})\otimes_{E}{A},

    d2(1w1)=αmα1Supp(g)i=1mc(αmα1)αmαi+1αiαi1α1,d_{2}(1\otimes w\otimes 1)=\sum_{\alpha_{m}\cdots\alpha_{1}\in\mathrm{Supp}(g)}\sum_{i=1}^{m}c(\alpha_{m}\cdots\alpha_{1})\alpha_{m}\cdots\alpha_{i+1}\otimes\alpha_{i}\otimes\alpha_{i-1}\cdots\alpha_{1},

    where g𝒢g\in\mathcal{G} such that g=w+pQ0;pwc(p)pg=w+\sum_{p\in Q_{\geq 0};\ p\neq w}c(p)p with Tip(g)=w\mathrm{Tip}(g)=w and 0c(p)k0\neq c(p)\in k. (Note that c(w)=1c(w)=1 and l(w)2l(w)\geq 2.)

Proof.

By the identifications at the end of Section 2.4, we just need to check the differentials in this lemma. Obviously, d0=μAd_{0}=\mu_{A} which inherits from the reduced bar resolution of AA, is given by the multiplication of AA. By the definition of the Morse matching \mathcal{M} in the weighted quiver G(B)G(B_{*}) (cf. Section 2.4), there are no dotted arrows starting from W(1)=Q0W^{(-1)}=Q_{0} in the new weighted quiver G(B)G_{\mathcal{M}}(B_{*}), where B:=B(A)B_{*}:=B(A) is the reduced bar resolution of AA. Thus for [α]W(0)[\alpha]\in W^{(0)} with αQ1\alpha\in Q_{1}, the zigzag paths from W(0)=Q1W^{(0)}=Q_{1} to W(1)=Q0W^{(-1)}=Q_{0} can be given by