12.10.2014 Views

The cohomology structure of string algebras

The cohomology structure of string algebras

The cohomology structure of string algebras

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>The</strong> <strong>cohomology</strong> <strong>structure</strong> <strong>of</strong> <strong>string</strong> <strong>algebras</strong><br />

Juan Carlos Bustamante<br />

Abstract. We show that the graded commutative ring <strong>structure</strong> <strong>of</strong> the Hochschild<br />

<strong>cohomology</strong> HH ∗ (A) is trivial in case A is a triangular quadratic <strong>string</strong><br />

algebra. Moreover, in case A is gentle, the Lie algebra <strong>structure</strong> on HH ∗ (A) is<br />

also trivial.<br />

Introduction<br />

Let k be a field, and A be a finite dimensional k−algebra. In this situation,<br />

the Hochschild <strong>cohomology</strong> groups HH i (A, M) with coefficients in some<br />

A − A−bimodule M can be identified with the groups Ext i A−A(A, M). In case<br />

AM A = A A A , we simply write HH i (A). <strong>The</strong> sum HH ∗ (A) = ∐ i≥0 HHi (A) is a<br />

graded commutative ring under the Yoneda product, which coincides with a cupproduct<br />

∪. Beside this, there is another product, namely the bracket product [−, −]<br />

which makes <strong>of</strong> HH ∗ (A) a graded Lie algebra. Moreover, these two <strong>structure</strong>s are<br />

related so that HH ∗ (A) is in fact a Gerstenhaber algebra [9].<br />

In general, very little information is known about these <strong>structure</strong>s. As Green<br />

and Solberg point out in [11], "the ring <strong>structure</strong> <strong>of</strong> HH ∗ (A) has <strong>of</strong>ten been observed<br />

to be trivial". Moreover, in [7], Cibils showed that if A = kQ/F 2 , where F is the<br />

two sided ideal <strong>of</strong> kQ generated dy the arrows, and Q is not an oriented cycle, then<br />

the ring <strong>structure</strong> <strong>of</strong> HH ∗ (A) is trivial. Although "one knows that for many self<br />

injective rings there are non zero products in HH ∗ (A)", there are rather few known<br />

examples <strong>of</strong> <strong>algebras</strong> having finite global dimension such that the ring <strong>structure</strong> <strong>of</strong><br />

HH ∗ (A) is not trivial. This leads us to state the following:<br />

Conjecture. Let A = kQ/I be a monomial triangular algebra. <strong>The</strong>n the ring<br />

<strong>structure</strong> <strong>of</strong> HH ∗ (A) is trivial.<br />

In case A = kQ/I is a monomial algebra, one has a precise description <strong>of</strong> a<br />

minimal resolution <strong>of</strong> A A A . <strong>The</strong> A − A−projective bimodules appearing in this<br />

resolution are described in terms <strong>of</strong> paths <strong>of</strong> Q (see [3]). Green, Snashall and<br />

Solberg use this resolution to achieve a deep study <strong>of</strong> Hochschild <strong>cohomology</strong> ring,<br />

modulo nilpotent elements <strong>of</strong> a monomial algebra in [10]. In the general case, this<br />

description leads to hard combinatorial computations. However, in case A is a<br />

monomial quadratic algebra, the minimal resolution is particularly easy to handle.<br />

Among monomial <strong>algebras</strong>, the class <strong>of</strong> <strong>string</strong> <strong>algebras</strong>, which are always tame, is<br />

Subject classification: 16E40. Keywords and phrases: Hochschild <strong>cohomology</strong> rings, <strong>string</strong><br />

<strong>algebras</strong>.


2 J. C. BUSTAMANTE<br />

particularly well understood, at least from the representation theoretic point <strong>of</strong> view<br />

[5]. We study the <strong>cohomology</strong> <strong>structure</strong> <strong>of</strong> triangular quadratic <strong>string</strong> <strong>algebras</strong>.<br />

After recalling some notions concerning the products in HH ∗ (A) in section 1,<br />

we establish some technical preliminary results in section 2. Finally, section 3 is<br />

devoted to state and show the main result <strong>of</strong> this work, which proves the conjecture<br />

below in the case <strong>of</strong> triangular quadratic <strong>string</strong> <strong>algebras</strong>.<br />

<strong>The</strong>orem 3.1. Let A = kQ/I be a triangular quadratic <strong>string</strong> algebra, n > 0<br />

and m > 0. <strong>The</strong>n HH n (A) ∪ HH m (A) = 0.<br />

In addition, we obtain a similar result concerning the Lie algebra <strong>structure</strong> for<br />

a particular case <strong>of</strong> <strong>string</strong> <strong>algebras</strong>, the so-called gentle <strong>algebras</strong>.<br />

<strong>The</strong>orem 3.2. Let A = kQ/I be a triangular gentle algebra, n > 1 and m > 1.<br />

<strong>The</strong>n [HH n (A), HH m (A)] = 0.<br />

1. Preliminaries<br />

1.1. Algebras. Let Q = (Q 0 , Q 1 , s, t) be a finite quiver (see [4]), and k be a<br />

commutative field. We consider <strong>algebras</strong> <strong>of</strong> the form kQ/I, where I an admissible<br />

ideal <strong>of</strong> the path algebra kQ. This includes, for instance, all basic connected finite<br />

dimensional <strong>algebras</strong> over algebraically closed fields (see [4]).<br />

While we briefly recall some particular concepts concerning bound quivers and<br />

<strong>algebras</strong>, we refer the reader to [4], for instance, for unexplained notions. <strong>The</strong><br />

composition <strong>of</strong> two arrows α 1 : s(α 1 ) t(α 1 ), and α 2 : s(α 2 ) t(α 2 ) such that<br />

s(α 2 ) = t(α 1 ) is the path α 1 α 2 , and will be denoted by α 1 α 2 : s(α 1 ) t(α 2 ).<br />

If Q has no oriented cycles, then A is said to be a triangular algebra. A two sided<br />

ideal I kQ generated by a set <strong>of</strong> paths in Q is said to be monomial. Moreover<br />

if the generators <strong>of</strong> an admissible ideal I are linear combinations paths <strong>of</strong> length<br />

2, then I is said to be quadratic. An algebra A = kQ/I is said to be monomial,<br />

or quadratic, depending on whether I is monomial or quadratic. In the remaining<br />

part <strong>of</strong> this paper, all <strong>algebras</strong> are triangular, quadratic and monomial.<br />

Among monomial <strong>algebras</strong>, the so-called <strong>string</strong> <strong>algebras</strong> [5] are particularly well<br />

understood, at least from the representation theoretic point <strong>of</strong> view. Recall from<br />

[5] that an algebra A = kQ/I is said to be a <strong>string</strong> algebra if (Q, I) satisfies the<br />

following conditions :<br />

(S1) I is a monomial ideal,<br />

(S2) Each vertex <strong>of</strong> Q is the source and the target <strong>of</strong> at most two arrows, and<br />

(S3) For an arrow α in Q there exists at most one arrow β and at most one<br />

arrow γ such that αβ ∉ I and γα ∉ I.<br />

A <strong>string</strong> algebra A = kQ/I is called a gentle algebra [2] if in addition (Q, I)<br />

satisfies :<br />

(G1) For an arrow α in Q there exists at most one arrow β and at most one<br />

arrow γ such that αβ ∈ I and γα ∈ I,<br />

(G2) I is quadratic.<br />

1.2. Hochschild <strong>cohomology</strong>. Given an algebra A over a field k, the Hochschild<br />

<strong>cohomology</strong> groups <strong>of</strong> A with coefficients in some A−A−bimodule M, denoted<br />

by HH i (A, M), are the groups Ext i A e(A, M) where Ae = A ⊗ k A op is the enveloping<br />

algebra <strong>of</strong> A. In case M is the A − A−bimodule A A A , we simply denote them


THE COHOMOLOGY STRUCTURE OF STRING ALGEBRAS 3<br />

by HH i (A). We refer the reader to [12], for instance, for general results about<br />

Hochschild (co)-homology <strong>of</strong> <strong>algebras</strong>.<br />

<strong>The</strong> lower <strong>cohomology</strong> groups, that is those for 0 ≤ i ≤ 2, have interpretations<br />

and give information about, for instance, the simple-connectedness or the rigidity<br />

properties <strong>of</strong> A (see [1, 8], for instance).<br />

Beside this, the sum HH ∗ (A) = ∐ i≥0 HHi (A) has an additional <strong>structure</strong> given<br />

by two products, namely the cup-product ∪, and the bracket [−, −], which are<br />

defined in [9] at the cochain level. <strong>The</strong> definitions are given in 1.5.<br />

1.3. Resolutions. From [6], we have a projective resolution <strong>of</strong> A over A e<br />

which is smaller that the standard Bar resolution. Let E be the (semi-simple)<br />

subalgebra <strong>of</strong> A generated by Q 0 . In the remaining part <strong>of</strong> this note, tensor products<br />

will be taken over E, unless it is explicitly otherwise stated. Note that, as E −<br />

E−bimodules, we have A ≃ E ⊕ radA. Let radA ⊗n denote the n th tensor power<br />

<strong>of</strong> radA with itself. With these notations, one has a projective resolution <strong>of</strong> A as<br />

A − A−bimodule, which we denote by K<br />

rad • (A)<br />

. . . A ⊗ radA ⊗n ⊗ A bn−1 A ⊗ radA ⊗n−1 ⊗ A bn−2 . . .<br />

A ⊗ radA ⊗ A<br />

where ɛ is the multiplication <strong>of</strong> A, and<br />

b 0<br />

A ⊗ A<br />

b n−1 (1 ⊗ r 1 ⊗ · · · ⊗ r n ⊗ 1) = r 1 ⊗ r 2 ⊗ · · · ⊗ r n ⊗ 1<br />

+<br />

ɛ<br />

A 0<br />

n−1<br />

∑<br />

(−1) j 1 ⊗ r 1 ⊗ · · · ⊗ r j r j+1 ⊗ · · · ⊗ r n ⊗ 1<br />

j=1<br />

+ (−1) n 1 ⊗ r 1 ⊗ · · · ⊗ r n .<br />

Remark. Note that since the ideal I is assumed to be monomial, radA is<br />

generated, as E − E−bimodule, by classes p = p + I <strong>of</strong> paths <strong>of</strong> Q. Moreover,<br />

since the tensor products are taken over E, then A ⊗ radA ⊗n ⊗ A is generated, as<br />

A − A−bimodule, by elements <strong>of</strong> the form 1 ⊗ p 1 ⊗ · · · ⊗ p n ⊗ 1 where p i are paths<br />

<strong>of</strong> Q such that the ending point <strong>of</strong> p i is the starting point <strong>of</strong> p i+1 , for each i such<br />

that 1 ≤ i < n.<br />

Keeping in mind that I is quadratic, the minimal resolution <strong>of</strong> Bardzell [3]<br />

has the following description: Let Γ 0 = Q 0 , Γ 1 = Q 1 , and for n ≥ 2 let Γ n =<br />

{α 1 α 2 · · · α n | α i α i+1 ∈ I, for 1 ≤ i < n}. For n ≥ 0, we let kΓ n be the E −<br />

E−bimodule generated by Γ n . With these notations, we have a minimal projective<br />

resolution <strong>of</strong> A as A − A−module, which we denote by K • min (A):<br />

. . . A ⊗ kΓ n ⊗ A δn−1 A ⊗ kΓ n−1 ⊗ A δn−2 . . .<br />

A ⊗ kΓ 1 ⊗ A δ0 A ⊗ kΓ 0 ⊗ A<br />

ɛ<br />

A 0<br />

where, again, ɛ is the composition <strong>of</strong> the isomorphism A ⊗ kΓ 0 ⊗ A ≃ A ⊗ A with<br />

the multiplication <strong>of</strong> A, and, given 1 ⊗ α 1 · · · α n ⊗ 1 ∈ A ⊗ kΓ n ⊗ A we have<br />

δ n−1 (1 ⊗ α 1 · · · α n ⊗ 1) = α 1 ⊗ α 2 · · · α n ⊗ 1 + (−1) n 1 ⊗ α 1 · · · α n−1 ⊗ α n .<br />

In order to compute the Hochschild <strong>cohomology</strong> groups <strong>of</strong> A, we apply the functor<br />

Hom A e(−, A) to the resolutions K min • (A) and K• rad<br />

(A). To avoid cumbersome<br />

notations we will write Crad n (A, A) instead <strong>of</strong> Hom A e(A ⊗ radA⊗n ⊗ A, A) which in


4 J. C. BUSTAMANTE<br />

addition we identify to Hom E e(radA ⊗n , A). In a similar way we define Cmin n (A, A),<br />

and, moreover, the obtained differentials will be denoted by b • , and δ • .<br />

<strong>The</strong> products in the <strong>cohomology</strong> HH ∗ (A) are induced by products defined using<br />

the Bar resolution <strong>of</strong> A (see [9]). Keeping in mind that our <strong>algebras</strong> are assumed<br />

to be triangular (so that the elements <strong>of</strong> Cmin n (A, A) take values in radA), these<br />

products are easily carried to K<br />

rad • (A) (see section 1.5). However, the spaces involved<br />

in that resolution are still "too big" to work with. We wish to carry these<br />

products from K<br />

rad • (A) to K• min (A). In order to do so, we need explicit morphisms<br />

µ • : K min • (A) K<br />

rad • (A), and ω • : K<br />

rad • (A) K min • (A) (compare with<br />

[13] p.736).<br />

Define a morphism <strong>of</strong> A−A−bimodules µ n : A ⊗ kΓ n ⊗ A A ⊗ radA ⊗n ⊗ A<br />

by the rule µ n (1⊗α 1 · · · α n ⊗1) = 1⊗α 1 ⊗· · ·⊗α n ⊗1. A straightforward computation<br />

shows that µ • = (µ n ) n≥0 is a morphism <strong>of</strong> complexes, and that each µ n splits.<br />

On the other hand, an element 1 ⊗ p 1 ⊗ · · · ⊗ p n ⊗ 1 <strong>of</strong> A ⊗ radA ⊗n ⊗ A, can<br />

always be written as 1 ⊗ p ′ 1α 1 ⊗ p 2 ⊗ · · · ⊗ α n p ′ n ⊗ 1 with α 1 , α n arrows in Q. Define<br />

a map ω n : A ⊗ radA ⊗n ⊗ A A ⊗ kΓ n ⊗ A by the rule<br />

{<br />

ω n (1⊗p ′ 1α 1 ⊗· · ·⊗α n p ′ p<br />

′<br />

n⊗1) = 1 ⊗ α 1 p 2 · · · p n−1 α n ⊗ p ′ n if α 1 p 2 · · · p n−1 α n ∈ Γ n<br />

0 otherwise.<br />

Again, ω • = (ω n ) n≥0 defines a morphism <strong>of</strong> complexes, which is an inverse for µ • .<br />

Summarizing what precedes, we obtain the following lemma.<br />

1.4. Lemma. With the above notations,<br />

a) µ • : K min • (A) K<br />

rad • (A), and ω • : K<br />

rad • (A) K min • (A) are morphisms<br />

<strong>of</strong> complexes, and<br />

b) ω • µ • = id, and thus, since the involved complexes have the same homology,<br />

these morphisms are quasi-isomorphisms.<br />

1.5. Products in <strong>cohomology</strong>. Following [9], given f ∈ Crad n (A, A), g ∈<br />

Crad m (A, A), and i ∈ {1, . . . , n}, define the element f ◦ i g ∈ C n+m−1<br />

rad<br />

(A, A) by the<br />

rule<br />

f◦ i g(r 1 ⊗· · ·⊗r n+m−1 ) = f(r 1 ⊗· · ·⊗r i−1 ⊗g(r i ⊗· · ·⊗r i+m−1 )⊗r i+m ⊗· · ·⊗r n+m−1 ).<br />

<strong>The</strong> composition product f ◦ g is then defined as ∑ n<br />

i=1 (−1)(i−1)(m−1) f ◦ i g. Let<br />

f ◦ g = 0 in case n = 0, and define the bracket<br />

[f, g] = f ◦ g − (−1) (n−1)(m−1) g ◦ f.<br />

On the other hand, the cup-product f ∪ g ∈ C n+m (A, A) is defined by the rule<br />

rad<br />

□<br />

(f ∪ g)(r 1 ⊗ · · · ⊗ r n+m ) = f(r 1 ⊗ · · · ⊗ r n )g(r n+1 ⊗ · · · ⊗ r n+m ).<br />

Recall that a Gerstenhaber algebra is a graded k−vector space A endowed<br />

with a product which makes A into a graded commutative algebra, and a bracket<br />

[−, −] <strong>of</strong> degree −1 that makes A into a graded Lie algebra, and such that [x, yz] =<br />

[x, y]z + (−1) (|x|−1)|y| y[x, z], that is, a graded analogous <strong>of</strong> a Poisson algebra.<br />

<strong>The</strong> cup product ∪ and the bracket [−, −] define products (still denoted ∪<br />

and [−, −]) in the Hochschild <strong>cohomology</strong> ∐ i≥0 HHi (A), which becomes then a<br />

Gerstenhaber algebra (see [9]).


THE COHOMOLOGY STRUCTURE OF STRING ALGEBRAS 5<br />

Using µ • and ω • we can define analogous products in C min • (A, A). We study<br />

these products later.<br />

2. Preparatory lemmata<br />

Recall that all our <strong>algebras</strong> are assumed to be triangular, monomial and quadratic.<br />

<strong>The</strong> spaces Cmin n (A, A) = Hom E e(kΓ n, A) have natural bases B n that<br />

we identify to the sets {(α 1 · · · α n , p)| α 1 · · · α n ∈ Γ n , p is a path with s(p) =<br />

s(α 1 ), t(p) = t(α n ), p ∉ I}. More precisely the map <strong>of</strong> E − E−bimodules<br />

f : kΓ n<br />

A corresponding to (α 1 · · · α n , p) is defined by<br />

{ p if γ1 · · · γ<br />

f(γ 1 · · · γ n ) =<br />

n = α 1 · · · α n ,<br />

0 otherwise.<br />

Given a path p in Q, denote by < p > the two sided ideal <strong>of</strong> kQ generated by<br />

p. We distinguish three different kinds <strong>of</strong> basis elements in Cmin n (A, A), which yield<br />

three subspaces C−, n C+, n and C0<br />

n <strong>of</strong> Cmin n (A, A):<br />

(1) C− n is generated by elements <strong>of</strong> the form (α 1 · · · α n , α 1 p),<br />

(2) C0 n is generated by elements <strong>of</strong> the form (α 1 · · · α n , w) such that w ∉<<br />

α 1 >, w ∉< α n >, and<br />

(3) C+ n is generated by elements <strong>of</strong> the form (α 1 · · · α n , qα n ) such that q ∉<<br />

α 1 >.<br />

Clearly, as vector spaces we have C n min (A, A) = Cn − ∐ C n 0 ∐ C n +. Moreover, let<br />

B − , B 0 , and B + be the natural bases <strong>of</strong> C n −, C n +, and C n 0 .<br />

Consider an element f = (α 1 · · · α n , α 1 p), ∈ B n −. First <strong>of</strong> all, since α 1 α 2 ∈ I,<br />

then p ∉< α 2 >, and, since A is assumed to be triangular, then p ∉< α 1 >. <strong>The</strong><br />

following figure illustrates this situation:<br />

α<br />

e<br />

1 α<br />

0<br />

e<br />

2<br />

1<br />

<br />

<br />

α<br />

· · · e n α<br />

n−1 n+1<br />

<br />

e n<br />

e n+1<br />

p<br />

Now, define f + : kΓ n<br />

A in the following way:<br />

{ (−1)<br />

f + (γ 1 · · · γ n ) =<br />

n+1 pα n+1 if γ 1 · · · γ n = α 2 · · · α n+1 ∈ Γ n ,<br />

0 otherwise.<br />

Analogously, given g = (α 1 · · · α n , qα n ) ∈ B n +, we define g − : kΓ n<br />

{<br />

α0 q if γ<br />

g − (γ 1 · · · γ n ) =<br />

1 · · · γ n = α 0 · · · α n−1 ∈ Γ n ,<br />

0 otherwise.<br />

It is easily seen that f + ∈ C n +, and g − ∈ C n −.<br />

A by<br />

2.1. Lemma. Let A = kQ/I be a <strong>string</strong> triangular algebra, f ∈ B n −, and<br />

g ∈ B n +.<br />

a) f − f + ∈ Imδ n−1 , and<br />

b) g − g − ∈ Imδ n−1 .<br />

Pro<strong>of</strong> : We only prove statement a). Let h = (α 2 · · · α n , p) ∈ C n−1<br />

min (A, A). We<br />

show that f + = f − δ n−1 h, indeed:


6 J. C. BUSTAMANTE<br />

δ n−1 h(α 1 · · · α n ) = α 1 h(α 2 · · · α n ) + (−1) n h(α 1 · · · α n−1 )α n<br />

= α 1 p<br />

= f(α 1 · · · α n ),<br />

thus (f − δ n−1 h)(α 1 · · · α n ) = 0 = f + (α 1 · · · α n ).<br />

Now let p = βp ′ with β an arrow. Since α 1 p /∈ I, then α 1 β ∉ I. But A is<br />

a <strong>string</strong> algebra, thus condition (S3) ensures that for every arrow α ′ 1 such that<br />

t(α 1 ) = t(α ′ 1) we have α ′ 1β ∈ I, thus α ′ 1p ∈ I.<br />

α<br />

x<br />

1 α<br />

1<br />

x<br />

2 α<br />

2<br />

x 3<br />

· · · x<br />

n−1<br />

n−1 x n<br />

<br />

α ′ 1<br />

β <br />

y<br />

<br />

x ′ 1<br />

Assume α ′ 1 is such that α ′ 1α 2 · · · α n ∈ Γ n . In particular α ′ 1α 2 ∈ I, so that<br />

δ n−1 h(α ′ 1α 2 · · · α n ) = α ′ 1h(α 2 · · · α n ) + (−1) n h(α ′ 1 · · · α n−1 )α n<br />

= α ′ 1p<br />

= 0.<br />

Thus, again, (f − δ n−1 h)(α ′ 1 · · · α n ) = 0 = f + (α ′ 1 · · · α n ).<br />

On the other hand, assume there exists an arrow α n+1 such that α 2 · · · α n+1 ∈<br />

Γ n . <strong>The</strong>n we have<br />

δ n−1 h(α 2 · · · α n+1 ) = α 2 h(α 3 · · · α n+1 ) + (−1) n h(α 2 · · · α n )<br />

= (−1) n h(α 2 · · · α n )α n+1<br />

p ′<br />

and hence<br />

= (−1) n pα n+1<br />

(f − δ n−1 h)(α 2 · · · α n+1 ) = 0 − (−1) n pα n+1<br />

= f + (α 2 · · · α n+1 ).<br />

Finally, note that f, f + , and h vanish on any other element γ 1 · · · γ n <strong>of</strong> Γ n .<br />

According to the decomposition C n min (A, A) = Bn −<br />

∐ B<br />

n<br />

0<br />

∐ B<br />

n<br />

+ , given an element<br />

φ ∈ C n min (A, A) one can write φ = f + h + g, where f ∈ Bn −, h ∈ B n 0 , and g ∈ B n +.<br />

Define then<br />

φ ≤ = f + + h + g and φ ≥ = f + h + g − .<br />

Remark. It follows from the preceding lemma that φ−φ ≤ , and φ−φ ≥ belong<br />

to Imδ n−1 . This will be useful later.<br />

2.2. Lemma. Let A be a triangular <strong>string</strong> algebra, φ ∈ Cmin n (A, A), and<br />

α 1 · · · α n ∈ Γ n . <strong>The</strong>n<br />

a) φ ≤ (α 1 · · · α n ) = (W + Qα n ) + I where W is a linear combination <strong>of</strong> paths<br />

none <strong>of</strong> which belongs to < α 1 > nor to < α n >, and Q is a linear<br />

combination <strong>of</strong> paths none <strong>of</strong> which belongs to < α 1 >, and<br />

b) φ ≥ (α 1 · · · α n ) = (α 1 P + W ′ ) + I, where W ′ is a linear combination <strong>of</strong><br />

paths none <strong>of</strong> which belongs to < α 1 > nor to < α n >, and P is a linear<br />

combination <strong>of</strong> paths none <strong>of</strong> which belongs to < α n >.<br />


THE COHOMOLOGY STRUCTURE OF STRING ALGEBRAS 7<br />

Pro<strong>of</strong> : By construction we have φ ≤ ∈ B n 0<br />

∐ B<br />

n<br />

+ , and φ ≥ ∈ B n −<br />

∐ B<br />

n<br />

0 . □<br />

2.3. Lemma. Let φ ∈ Kerδ n , and α 1 · · · α n ∈ Γ n such that φ(α 1 · · · α n ) ≠ 0.<br />

a) If there exists α n+1 such that α 1 · · · α n+1 ∈ Γ n+1 then φ ≤ (α 1 · · · α n )α n+1 =<br />

0,<br />

b) If there exists α 0 such that α 0 · · · α n ∈ Γ n+1 then α 0 φ ≥ (α 1 · · · α n ) = 0<br />

Pro<strong>of</strong> : We only prove a). Let φ ∈ Kerδ n . <strong>The</strong>n, since φ − φ ≤ ∈ Imδ n−1 , we have<br />

φ ≤ ∈ Kerδ n , thus<br />

0 = δ n φ ≤ (α 1 · · · α n α n+1 )<br />

= α 1 φ ≤ (α 2 · · · α n+1 ) + (−1) n+1 φ ≤ (α 1 · · · α n )α n+1 .<br />

<strong>The</strong> statement follows from the fact that paths are linearly independent, and from<br />

the preceding lemma.<br />

□<br />

3. <strong>The</strong> <strong>cohomology</strong> <strong>structure</strong><br />

<strong>The</strong> morphisms µ • : K min • (A) K<br />

rad • (A), and ω • : K<br />

rad • (A) K min • (A) allow<br />

us to carry the products defined in C<br />

rad • (A, A) to C• min (A, A). In this way we<br />

obtain a cup product and a bracket, which we still denote ∪ and [−, −]. More<br />

precisely, applying the functor Hom A e(−, A) and making the identifications <strong>of</strong> section<br />

1.5, we obtain morphisms <strong>of</strong> complexes µ • : C min • (A, A) C<br />

rad • (A, A), and<br />

ω • : C<br />

rad • (A, A) C min • (A, A). Given f ∈ Cn min (A, A), g ∈ Cm min (A, A), define<br />

f ∪ g ∈ C n+m<br />

min (A, A) as<br />

f ∪ g = µ n+m (ω n f ∪ ω m f).<br />

Thus, given an element α 1 · · · α n β 1 · · · β m ∈ Γ n+m , we have<br />

f ∪ g(α 1 · · · α n β 1 · · · β m ) = f(α 1 · · · α n )g(β 1 · · · β m ).<br />

As usual, we have δ n+m (f ∪ g) = δ n f ∪ g + (−1) m f ∪ δ m g, so the product ∪<br />

defined in C min • (A, A) induces a product at the <strong>cohomology</strong> level. In fact, the latter<br />

coincides with the cup-product <strong>of</strong> section 1.5 and the Yoneda product. In what<br />

follows, we work with the product ∪ defined in C min • (A, A).<br />

3.1. <strong>The</strong>orem. Let A = kQ/I be a triangular quadratic <strong>string</strong> algebra, n > 0<br />

and m > 0. <strong>The</strong>n HH n (A) ∪ HH m (A) = 0.<br />

Pro<strong>of</strong> : Let φ ∈ HH n (A), and ψ ∈ HH m (A). We will show that φ ≤ ∪ ψ ≥ = 0<br />

in HH n+m (A). Assume the contrary is true. In particular there exist an element<br />

α 1 · · · α n β 1 · · · β m ∈ Γ n+m such that<br />

φ ≤ (α 1 · · · α n )ψ ≥ (β 1 · · · β m ) ≠ 0.<br />

With the notations <strong>of</strong> Lemma 2.2, this is equivalent to<br />

(1) (W + Qα n )(W ′ + β 1 P ) ∉ I.<br />

Now, since α n β 1 ∈ I, Lemma 2.3 gives<br />

(W + Qα n )β 1 ∈ I, and α 1 (W ′ + β 1 P ) ∈ I.<br />

Thus, equation (1) gives W W ′ ∉ I. Let w and w ′ be paths appearing in W and<br />

W ′ respectively, such that ww ′ ∉ I. Moreover, write w = uγ, w ′ = γ ′ u ′ with γ, γ ′<br />

arrows <strong>of</strong> Q. Since ww ′ ∉ I, we have that γγ ′ ∉ I. But then, since (Q, I) satisfies<br />

(S3), this implies α n γ ′ ∈ I, so that α 1 · · · α n γ ′ ∈ Γ n+1 and, again, Lemma 2.3 gives<br />

(W + Qα n )γ ′ ∈ I so that W γ ′ ∈ I, and, in particular wγ ′ ∈ I, a contradiction. □


8 J. C. BUSTAMANTE<br />

Remark. At this point, it is important to note that even if Lemmas 2.1, 2.2 and<br />

2.3 were stated assuming that A = kQ/I is a <strong>string</strong> triangular algebra, the condition<br />

(S2) in the definition <strong>of</strong> <strong>string</strong> <strong>algebras</strong> has never been used. Thus, the preceding<br />

theorem holds for every monomial quadratic, triangular algebra A = kQ/I such<br />

that (Q, I) satisfies (S3). This includes, for instance, the gentle <strong>algebras</strong>.<br />

We now turn our attention to the Lie algebra <strong>structure</strong> HH ∗ (A).<br />

Given φ ∈ Cmin n (A, A), ψ ∈ Cm min (A, A), and i ∈ {1, . . . , n} we define φ ◦ i ψ ∈<br />

(A, A) as<br />

φ ◦ i ψ = µ n+m−1 ((ω n φ) ◦ i (ω m ψ))<br />

C n+m−1<br />

min<br />

and from this, the bracket [−, −] is defined as in 1.5. In particular, and this is the<br />

crucial point for what follows, given α 1 · · · α n+m−1 ∈ Γ n+m−1 we have<br />

⎧<br />

⎨ φ(α 1 · · · α i−1 ψ(α i · · · α i+m−1 )α i+m · · · α n+m−1 ), if<br />

φ◦ i ψ(α 1 · · · α n+m−1 ) = α 1 · · · α i−1 ψ(α i · · · α i+m−1 )α i+m · · · α n+m−1 ∈ Γ n ,<br />

⎩<br />

0 otherwise.<br />

Again, one can verify that δ n+m−1 [f, g] = (−1) m−1 [δ n f, g] + [f, δ m g], so that<br />

[−, −] induces a bracket, which we still denote by [−, −] at the <strong>cohomology</strong> level<br />

(note that, by construction, this bracket coincides with that <strong>of</strong> 1.5).<br />

This leads us to the following result<br />

3.2. <strong>The</strong>orem. Let A = kQ/I be a triangular gentle algebra. <strong>The</strong>n, for n > 1<br />

and m > 1, we have [HH n (A), HH m (A)] = 0.<br />

Pro<strong>of</strong> : We show that, in fact, under the hypothesis, the products ◦ i are equal<br />

to zero at the cochain level. This follows immediately from the discussion above.<br />

Indeed, with that notations, if α 1 · · · α n+m−1 ∈ Γ n+m−1 we have, in particular that<br />

α i−1 α i ∈ I. But A being gentle, there is no other arrow β in Q with s(β) = t(α i−1 )<br />

such that α i−1 β ∈ I.<br />

□<br />

<strong>The</strong> following example shows that the previous result can not be extended to<br />

<strong>string</strong> <strong>algebras</strong> which are not gentle.<br />

3.3. Example. Let A = kQ/I where Q is the quiver<br />

3<br />

α 2<br />

<br />

2<br />

α 1 β<br />

<br />

1<br />

γ<br />

α 3<br />

4<br />

α 4<br />

5<br />

and I is the ideal generated by paths <strong>of</strong> length 2. This is a <strong>string</strong> algebra which is<br />

not gentle. From theorem 3.1 in [7] we get<br />

⎧<br />

⎨ 1 if i ∈ {0, 2, 3, 4},<br />

dim k HH i (A) = 2 if i = 1,<br />

⎩<br />

0 otherwise.<br />

Keeping in mind the identifications <strong>of</strong> section 2, the generators <strong>of</strong> HH i (A) are<br />

the elements corresponding to (β, ¯β), and (γ, ¯γ) for i = 1; (α 2 α 3 , ¯β) for i = 2;<br />

(α 1 βα 4 , ¯γ), for i = 3; and (α 1 α 2 α 3 α 4 , ¯γ) for i = 4. A straightforward computation<br />

shows that [HH n (A), HH m (A)] = HH n+m−1 (A).


THE COHOMOLOGY STRUCTURE OF STRING ALGEBRAS 9<br />

<strong>The</strong> first Hochschild <strong>cohomology</strong> group <strong>of</strong> an algebra A is by its own right a<br />

Lie algebra. In case the algebra A is monomial, this <strong>structure</strong> has been studied in<br />

[13]. <strong>The</strong> following result gives information about the role played by HH 1 (A) in<br />

the whole Lie algebra HH ∗ (A) in our context.<br />

3.4. Proposition. Let A = kQ/I be a triangular monomial and quadratic<br />

algebra. <strong>The</strong>n [HH n (A), HH 1 (A)] = HH n (A), whenever n > 1.<br />

Pro<strong>of</strong> : In fact, for every f ∈ Cmin n (A, A) there exists g ∈ C1 min (A, A) such that<br />

[f, g] = f, and g ≠ 0 in HH 1 (A). Clearly it is enough to consider an arbitrary basis<br />

element f ∈ Cmin n (A, A). Let f be such an element, corresponding to (α 1 · · · α n , p).<br />

Define g ∈ Cmin 1 (A, A) by {<br />

α1 if γ = α<br />

g(γ) =<br />

1 ,<br />

0 otherwise<br />

It is straightforward to verify that f ◦ g = f ◦ 1 g, and, since we assume A triangular<br />

and n > 1, that g ◦ f = 0 so that [f, g] = f. Moreover, direct computations show<br />

that ḡ ≠ 0 in HH 1 (A).<br />

□<br />

Acknowledgements. <strong>The</strong> author gratefully thanks pr<strong>of</strong>essors E.N. Marcos,<br />

for several interesting discussions and comments, as well as pr<strong>of</strong>essors E.N. Marcos<br />

and I. Assem for a carefully reading <strong>of</strong> preliminar versions <strong>of</strong> this work. This work<br />

was done while the author had post-doctoral fellowship at the University <strong>of</strong> São<br />

Paulo, Brazil. He gratefully acknowledges the University for hospitality during his<br />

stay there, as well as financial support from F.A.P.E.S.P.<br />

References<br />

[1] I. Assem and J.A. de la Peña. <strong>The</strong> fundamental groups <strong>of</strong> a triangular algebra. Comm.<br />

Algebra, 24(1):187–208, 1996.<br />

[2] I. Assem and A. Skowroński. Iterated tilted <strong>algebras</strong> <strong>of</strong> type Ã. Math. Zeitschrift, 195(2):269–<br />

290, 1987.<br />

[3] M. J. Bardzell. <strong>The</strong> alternating syzygy beheavior <strong>of</strong> monomial <strong>algebras</strong>. J. Algebra, 188:69–89,<br />

1997.<br />

[4] K. Bongartz and P. Gabriel. Covering spaces in representation theory. Invent. Math,<br />

65(3):331–378, 1981-1982.<br />

[5] M.C.R. Butler and C.M. Ringel. Auslander-Reiten sequences with few middle terms and<br />

applications to <strong>string</strong> <strong>algebras</strong>. Comm. Algebra, 15(1-2):145–179, 1987.<br />

[6] C. Cibils. Cohomology <strong>of</strong> incidence <strong>algebras</strong> and simplicial complexes. J. Pure Appl. Algebra,<br />

56:221–232, 1989.<br />

[7] C. Cibils. Hochschild coohomology <strong>of</strong> radical square zero <strong>algebras</strong>. In I. Reiten, S. Smalø, and<br />

Ø. Solberg, editors, Algebras and Modules II, number 24 in CMS Conf. proceedings, pages<br />

93–101, Geiranger, Norway, 1998.<br />

[8] M. Gerstenhaber. On the deformations <strong>of</strong> rings and <strong>algebras</strong>. Ann. Math. Stud., 79(2):59–103,<br />

1964.<br />

[9] M. Gerstenhaber. <strong>The</strong> Cohomology <strong>structure</strong> on an associative ring. Ann. <strong>of</strong> Math.,<br />

78(2):267–288, 1978.<br />

[10] E.L. Green, N. Snashall, and Ø. Solberg. <strong>The</strong> hochschild <strong>cohomology</strong> ring modulo nilpotence<br />

<strong>of</strong> a monomial algebra. www.arXiv:math/RT/0401446v1, 2004.<br />

[11] E.L. Green and Ø. Solberg. Hochschild <strong>cohomology</strong> rings and triangular rings. In Beijing<br />

Norm. Univ. Press, editor, Representations <strong>of</strong> algebra., volume I, pages 192–200, 2002.<br />

[12] D. Happel. Hochschild <strong>cohomology</strong> <strong>of</strong> finite-dimensional <strong>algebras</strong>, pages 108–126. Number<br />

1404 in Lecture Notes in Mathematics. Springer-Verlag, Berlin Heidelberg New York Tokyo,<br />

1989.<br />

[13] C. Strametz. <strong>The</strong> Lie algebra <strong>structure</strong> <strong>of</strong> the first Hochschild <strong>cohomology</strong> group for monomial<br />

<strong>algebras</strong>. C. R., Math., Acad. Sci. Paris, 334(9):733–738, 2002.


10 J. C. BUSTAMANTE<br />

Département de Mathématiques, Université de Sherbrooke, 2500 Boulevard de<br />

l’Université, Sherbrooke J1K 2R1, Québec, Canada.<br />

E-mail address: juan.carlos.bustamante@usherbrooke.ca

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!