Skip to content
Licensed Unlicensed Requires Authentication Published by De Gruyter May 3, 2018

On cohomological Hall algebras of quivers: Generators

  • Olivier Schiffmann and Eric Vasserot

Abstract

We study the cohomological Hall algebra Y of a Lagrangian substack Λ of the moduli stack of representations of the preprojective algebra of an arbitrary quiver Q, and their actions on the cohomology of Nakajima quiver varieties. We prove that Y is pure and we compute its Poincaré polynomials in terms of (nilpotent) Kac polynomials. We also provide a family of algebra generators. We conjecture that Y is equal, after a suitable extension of scalars, to the Yangian 𝕐 introduced by Maulik and Okounkov. As a corollary, we prove a variant of Okounkov’s conjecture, which is a generalization of the Kac conjecture relating the constant term of Kac polynomials to root multiplicities of Kac–Moody algebras.

A Appendix

A.1 Proof of Proposition 3.1

The proposition is proved in the literature under the assumption that Q does not carry oriented cycles (or edge loops). The following argument which works in full generality was explained to us by W. Crawley-Boevey. Let us begin by sketching the proof that Π is of homological dimension two. We may and will restrict ourselves to the case of a connected quiver. For any vertex iI we denote by ei the corresponding idempotent of Π. Consider the complex of (Π,Π)-bimodules

(A.1)

where all the tensor products are taken over , and where the maps are defined as follows: k is the multiplication map, g(aeh′′ehb)=ahehehb-aeh′′eh′′hb and

f(aeieib)=hΩ¯,h′′=iϵhae(h*)e(h*)′′h*b-hΩ¯,h=iϵhah*eh′′ehb,

where ϵh=1 if hΩ and ϵh=-1 if hΩ*. The fact that (A.1) is a complex is a direct consequence of the defining relations of Π. We claim that it is in addition exact. This exactness everywhere except for the leftmost term may be checked using standard arguments. Recall that we have assumed that Q is not of finite Dynkin type. Equipping Π with the -grading obtained by assigning degree 1 to any edge hΩ¯ we then have that Π is a Koszul algebra and that its Hilbert series

H(t)=(hi,j(t))i,jI,hi,j(t)=l0dim(ejΠ[l]ei)tl

is equal to HQ(t)=(Id+t(𝐐+𝐐t)+t2Id)-1, where 𝐐 is the adjacency matrix of Q, see [23, 14]. Observing that f,g,k are of respective degrees 1,1,0, we deduce that

dim(ker(f))=-H(t)+H(t)2-tH(t)(𝐐+𝐐t)H(t)+t2H(t)2=0

as wanted. This proves that Π is of homological dimension two. Next, let M, N be any finite-dimensional Π-modules. Tensoring (A.1) by M yields the projective resolution of M

0iΠeieiM𝑓hΩ¯Πeh′′ehM𝑔iΠeieiM𝑘M0.

Applying the functor Hom(N,), we obtain the following complex computing Ext(N,M):

0i(eiN)*eiMghΩ¯(eh′′N)*ehMfi(eiN)*eiM0.

It remains to observe that this complex is canonically isomorphic to the dual of the same complex in which the roles of M and N are exchanged.

A.2 Proof of Lemma 3.9

Being the fiber of the moment map 𝔤𝔩(v)2q𝔰𝔩(v), the set M(v) has all its irreducible components of dimension at least (2q-1)v2+1. Note also that M(v) is defined over and may be therefore reduced to any finite field. By the Lang–Weil theorem the irreducibility of M(v) will thus be a consequence of the estimate as l

#M(v)(𝔽l)l(2q-1)v2+1.

By [26, Theorem 5.1] the number of points of M(v)(𝔽l) is given by the generating series

(A.2)v0#M(v)(𝔽l)#GL(v)(𝔽l)lv,vzv=Exp(ll-1vAv(l)zv),

where Exp is the plethystic exponential and Av(l) is the Kac polynomial, see, e.g., [37, Section 2]. It is well known that Av(l) is a monic polynomial of degree 1-v,v=1+(g-1)v2. The coefficient of zv in (A.2) reads

(A.3)#M(v)(𝔽l)l(1-q)v2lv2i=1v(1-l-i)
=ll-1Av(l)+r21r!(ni,vi)i1iniilnilni-1Avi(lni),

where the sum ranges over all r-tuples ((n1,v1),,(nr,vr)) satisfying inivi=v. We claim that only the first term on the right-hand side of (A.3) contributes to the leading term. Indeed, ll-1Av(l)l1+(q-1)v2 while for any r2 and tuple (ni,vi)i we have

ilnilni-1Avi(lni)lini(1+(q-1)vi2)

and it is elementary to check that ini(1+(q-1)vi2)<1+(q-1)v2 whenever r2. It now follows by taking the leading term of (A.3) that #M(v)(𝔽l)l(2q-1)v2+1 as wanted.

Remark A.1.

One may alternatively use Beilinson and Drinfeld’s notion of a very good stack, see [1, Section 1.1.1.], and show that Rep(kQ,v)/G(v) is very good for any v and any Q=𝒬(i,q) with q>1.

A.3 Proof of Proposition 3.11

A simple representation S of Π~ will be called rigid if Ext1(S,S)={0}. Since

(dimS,dimS)=dim(Hom(S,S))-dim(Ext1(S,S))+dim(Ext2(S,S))=2-dim(Ext1(S,S)),

a simple representation S is rigid if and only if (dimS,dimS)=2. Alternatively, a representation S is rigid if and only if every other simple representation of the same dimension is isomorphic to S. On the other hand, if S is non-rigid, then there are infinitely many non-isomorphic simple representations of the same dimension (recall that we work over the field k=), and we have (dimS,dimS)0 (note that (u,u)2 for any dimension vectors u,u). For any dimension type d there exists a unique representation type τ=τ(d) of dimension type d in which all non-rigid simple representations occur with multiplicity one. The stratum 𝔐0(τ) is open and dense in 𝔐0((d)). Since π is proper, the set Im(π) is closed and the proposition will be proved once we show that for any d we have 𝔐0(τ)Im(π) as soon as there exists τ of dimension type d such that 𝔐0(τ)Im(π). By definition, the existence of such a τ means that there exists a stable representation MMs(v,w) of semisimple type τ. In other words, there exists a collection of simple Π~-modules S1,S2,,Sr such that #{i:dimSi=u}=du for any u, and an iterated extension

M1M2Mr,Mi/Mi-1Si

which is stable. Note that we automatically have dimS1I×{w} and dimSiI×{0} for any i>1. Our aim is then to prove, under the above hypothesis, the existence of a similar iterated extension N1N2Nr which is stable, for which Ti:=Ni/Ni-1 is simple, satisfies dimTi=dimSi and in which all non-rigid simple factors Ti are non-isomorphic. We will use the following simple observation, which might be of independent interest.

Lemma A.2.

Let N1N2Nr be a filtration with simple factors Ti:=Ni/Ni-1 such that dimT1NI×{0} while dimTiNI×{0} for any i>1. Let ξiExtΠ~1(Ti,Ni-1) be the associated elements. Then Nr is stable if and only if ξi0 for all i.

Proof.

The module Nr is stable if and only if it contains no submodule whose dimension vector belongs to I×{0}, and hence if and only if HomΠ~(Ti,Nj)={0} for all i,j. This in turn is equivalent to the condition that HomΠ~(Ti,Ni)={0} for all i. From the exact sequence

Hom(Ti,Ni-1)Hom(Ti,Ni)Hom(Ti,Ti)Ext1(Ti,Ni-1)

and the fact that ξi=(id) we see that Ni is stable if and only if ξi0 and Ni-1 is stable. The claim follows. ∎

Let us fix some w and prove the proposition by induction on v. Let us also fix a dimension type d such that udu=(v,w). Let us assume that Proposition 3.11 is proved for (w,v) for any dimension vector v<v. Let τ be a representation type of dimension type d¯, and of dimension (v,w). Let M1M2Mr be a filtration as above of representation type τ, and set Si=Mi/Mi-1. If r=1, there is nothing to prove, so we assume that r>1. There are two cases to consider.

Case 1: Sr is not rigid

By our induction hypothesis, there exists a filtration

N1N2Nr-1

such that Ti:=Ni/Ni-1 is a simple Π~-module of dimension dimSi, Nr-1 is stable and all non-rigid simple factors Tj are non-isomorphic. Let us choose a simple object Tr of the same dimension as Si, but non-isomorphic to all the Tj for j<i. This is possible since there are infinitely many non-isomorphic simples of dimension dimSr. We have to show that Ext1(Tr,Nr-1){0}. Let i0i be the smallest index for which dimTi0=dimSr. Since (dimTr,dimTr)0 and (dimTr,dimS)=-dimExt1(Tr,S)0 for any S which is not isomorphic to Tr, we have

(dimTr,dimNi0-1)<0,(dimTr,dim(Nr-1/Ni0-1))0,

therefore (dimTr,dimNi-1)<0 and hence Ext1(Tr,Ni-1){0}. We choose any nonzero ξExt1(Tr,Ni-1) to define Nr. By Lemma A.2 the representation Nr is stable, and by construction it is of representation type τ.

Case 2: Sr is rigid

Set n=dimHom(Mr,Sr). We claim that there exists a filtration

M1M2Mr-1Mr=Mr

of Mr for which Mr-j/Mr-j-1Sr for j=0,1,,n-1 and

Hom(Mr-n,Sr)={0},dimExt1(Sr,Mr-n)n.

Indeed, let M be any Π~-module satisfying

Hom(Sr,M)={0}andHom(M,Sr){0}.

Applying Hom(,Sr) to any exact sequence

0KMSr0

yields a short exact sequence

0Hom(M,Sr)Hom(K,Sr)0

from which we get dimHom(K,Sr)=dimHom(M,Sr)-1. Similarly, applying Hom(Sr,), we get the short exact sequence

0Hom(M,Sr)Hom(K,Sr)0

from which we deduce that dimExt1(Sr,K)=dimExt1(Sr,M)+1. Iterating this process n times starting with M, we get a filtration of Mr with the required properties. The representation type ν of Mr-n is obtained from τ by decreasing the multiplicity of Sr by n. By the induction hypothesis, there exists a stable representation Nr-n of type ν, obtained from τ by decreasing the multiplicity of Sk by n. Now, we have

(Sr,Nr-n)=(Sr,Mr-n)=dimHom(Sr,Mr-n)-dimExt1(Mr-n,Sr)+dimHom(Mr-n,Sr)=-dimExt1(Sr,Mr-n)-n

hence dimExt1(Sr,Nr-n)n. Reversing the process above, we may now consider n successive non-split extensions of Nr-n by Sr. The resulting module Nr is stable by Lemma A.2 and is of type τ as wanted.

The induction step is completed, and Proposition 3.11 is proved.

A.4 Kirwan surjectivity for local quiver varieties

In this subsection, we provide an independent proof of the Kirwan surjectivity for the fixed point quiver varieties used in this paper.

A.4.1 Local quiver varieties

We define a new quiver Q#=(I#,Ω#) defined as follows:

  1. I#={(i,l)}=I×,

  2. Ω#={(h,l):(h,l)(h′′,l-1)hΩ}{(h,l):(h,l)(h′′,l)hΩ*}.

For each dimension vectors v, w of Q# we define

R(v,w)#=Rep(kQ#,v)HomI#(W,V[-1])HomI#(V,W),

where we abbreviate V=kv, W=kw and (V[-1])i,l=Vi,l-1. The group G(v)×G(w) acts in the obvious way on R(v,w)#. The map

μ#:R(v,w)#HomI#(V,V[-1]),(x¯,a¯)[x,x*]+aa*

is G(v)×G(w)-equivariant. Let

M(v,w)#=μ#-1(0),
𝔐0(v,w)#=M(v,w)#//G(v).

The affine variety 𝔐0(v,w)# carries an action of the group T×G(w). Let s be the character of G(v) defined by

s(g)=i,ldet(gi,l)-1.

The local quiver variety associated with Q, v, w is the quasiprojective variety given by

𝔐s(v,w)#=Proj(nk[M(v,w)#]sn).

We may abbreviate 𝔐(v,w)#=𝔐s(v,w)#. There is a natural projective morphism

π:𝔐(v,w)#𝔐0(v,w)#.

A representation (x¯,a¯) in M(v,w)# is said to be semistable if and only if it does not admit any nonzero subrepresentation whose dimension vector belongs to I#×{0}. Let

Ms(v,w)#M(v,w)#

be the open subset consisting of the semistable points. Observe that this set is preserved by the T×G(v)×G(w)-action on R(v,w)#. The same arguments as in [28, 31] imply that 𝔐(v,w)# is the geometric quotient of Ms(v,w)# by G(v) and is smooth quasi-projective with an action of T×G(w). If Q# (or equivalently Q) has no oriented cycles, then 𝔐0(v,w)# is reduced to a single point and thus 𝔐(v,w)# is projective. However, this is not the case in general.

Note that the varieties 𝔐(v,w)# arise as a union of fixed point connected components for some suitable 𝔾m-actions on the Nakajima quiver varieties attached to the quiver Q, see Section 3.7.4. More precisely, fix dimension vectors u, y in I and consider the -action on 𝔐(u,y) in (4.2). We have

(A.4)𝔐(u,y)=v,w𝔐(v,w)#,

where the union ranges over the elements v, w in I# such that

lvi,l=ui,wi,l=δl,0yi.

This decomposition is compatible with the T-action.

A.4.2 The homology of local quiver varieties

For any vertex (i,l)I# the universal and the tautological T×G(w)-equivariant vector bundles over 𝔐(v,w)# are denoted by

𝒱i,l,𝒲i,l.

They are defined as in (5.22). Consider the subspace

Mv,w#M*T×G(w)(𝔐(v,w)#)

generated by the action of the Chern character components chk(𝒱i,l) and chk(𝒲i,l), for all possible i,l,k, on the fundamental class [𝔐(v,w)#].

Theorem A.3.

For any v, w we have Mv,w#=M*T×G(w)(M(v,w)#).

When Q has no oriented cycle, then Q# has no oriented cycle either, hence 𝔐(v,w)# is projective. In this case, the theorem follows from Nakajima’s resolution of the diagonal as in [30, Section 7.3]. We will prove the theorem by constructing a suitable compactification of the local quiver variety. A very similar construction has independently appeared in McGerty and Nevins’ recent proof of Kirwan surjectivity for (non-graded, non-local) quiver varieties, see [25].

Corollary A.4.

For any v,wNI#, the variety M(v,w)# is connected.

A.4.3 The quivers Q and Q¯

Consider a new quiver Q=(I,Ω) defined as follows:

  1. I={i1:iI}{i2:iI}I×{1,2},

  2. Ω={h:(h)1(h′′)2hΩ}{i:i1i2iI}.

Example

We have

The quiver Q is strongly bipartite: all arrows go from a vertex in I×{1} to one in I×{2}. In particular, it contains no oriented cycles. The moduli stacks of representations of Q and Q are related in the following fashion. For each vI we define vI such that

vi1=vi2=vi.

Let

Rep(kQ,v)Rep(kQ,v)

be the open subset consisting of the representations x such that xi is invertible for all i. Then the assignment x(x,g), where

xh=(xh′′)-1xh,gi=xi,

gives an isomorphism

Rep(kQ,v)Rep(kQ,v)×G(v),

hence an isomorphism of Artin stacks

Rep(kQ,v)/G(v)Rep(kQ,v)/G(v).

This allows us to view Rep(kQ,v)/G(v) as an open substack of Rep(kQ,v)/G(v).

We will now perform a similar construction for the double quiver. Let Q¯ be the double quiver of Q. For each v, w in I we define v, w in I such that

(A.5)vi1=vi2=vi,
wi1=wi,
wi2=0.

The group G(v) acts in a Hamiltonian fashion on

R(v,w)=Rep(kQ¯,v)HomI(w,v)HomI(v,w)

and we denote by μ:R(v,w)𝔤(v) the associated moment map. Let R(v,w) be the open subset of R(v,w) consisting of pairs (x¯,a¯) such that xi is an isomorphism for all i. We also put

M(v,w)=μ-1(0),
M(v,w)=R(v,w)M(v,w).

Now, let R(v,w) and M(v,w) be as in Section 3.7.1. Consider the map

ϕ:R(v,w)×G(v)R(v,w),((x¯,a¯),g)(x¯,a¯),

defined, for all hΩ and iI, by

(A.6)a¯i1=a¯i,
xh=gh′′xh,
xh*=xh*(gh′′)-1,
xi=gi,
xi*=-h′′=i(xi)-1xhxh*.

A direct computation gives:

Lemma A.5.

Given v,w let v,w be as in (A.5). Then the map ϕ gives an isomorphism M(v,w)×G(v)M(v,w). ∎

We will next consider appropriate stability conditions on R(v,w). Given a tuple of integers (θi), let θ be the character of G(v) defined by θ(g)=iIdet(gi)-θi. We set

𝔐θ(v,w)=Proj(nk[M(v,w)]θn),

which is a quasi-projective T×G(w)-variety equipped with a projective morphism

π:𝔐θ(v,w)𝔐0(v,w)=M(v,w)//G(v).

We can describe the set of closed points of 𝔐θ(v,w) explicitly. Set

θ=-iIθivi,W=iIWi,V=iIVi.

We say that an element (x¯,a¯) is θ-semistable if the following conditions are satisfied:

  1. For any (x¯,a¯)-stable I-graded subspace UV we have

    iIθidim(Ui)0.
  2. For any I-graded subspace UV such that UW is (x¯,a¯)-stable we have

    iIθidim(Ui)+θ0.

We will further say that (x¯,a¯) is stable if the above inequalities are strict for proper subspaces U. Let us denote by Rθ(v,w) the open subset of R(v,w) consisting of θ-semistable elements and by Mθ(v,w) its intersection with M(v,w). Then [32, Proposition 2.9] yields

𝔐θ(v,w)=Mθ(v,w)//G(v).

We also set

Mθ(v,w)=Mθ(v,w)R(v,w),
𝔐θ(v,w)=Mθ(v,w)//G(v).

We say that θ is generic if neither equations

iIθiui=0andiIθiui+θ=0

have integer solutions (ui) satisfying 0uivi other than the trivial solutions

(ui)=0or(ui)=(vi).

If θ is generic, then any semistable pair (x¯,a¯) is stable and in that case the map

Mθ(v,w)𝔐θ(v,w)

is a G(v)-torsor, hence the variety 𝔐θ(v,w) is smooth.

Lemma A.6.

Given v,w let v,w be as in (A.5). Let θZI be generic and such that

(A.7)θi10,θi20,θi1+θi2=1for all iI.

Then the map ϕ restricts to an isomorphism

Mθ(v,w)×G(v)Mθ(v,w).
Proof.

Fix a tuple z=(x¯,a¯) in M(v,w). We will prove that

zMθ(v,w)ϕ-1(z)Mθ(v,w)×G(v).

Let U be an I-graded subspace of V.

Let us first assume that U is z-stable and nonzero. Since xi is invertible, we have

dim(Ui1)dim(Ui2),iI.

If there exists jI such that dim(Uj1)<dim(Uj2), then

iIθidim(Ui)=iI[θi2(dim(Ui2)-dim(Ui1))+dim(Ui1)]θj2+iIvi<0

because θj20. Thus U is not destabilizing. On the other hand, if xi is an isomorphism Ui1Ui2 for all i, then we have

iIθidim(Ui)=idim(Ui1)>0

hence U is destabilizing.

Assume now that UW is z-stable and proper. As above, we have

dim(Ui1)dim(Ui2)

for all i. Thus

iIθidim(Ui)+θ=iI[θi2(dim(Ui2)-dim(Ui1))+dim(Ui1)]+θiIdim(Ui1)-vi<0

for otherwise dim(Ui1)=dim(Ui2)=vi for all i and U is not a proper subspace of V. Therefore U is not destabilizing.

By the above, we conclude that z is θ-semistable if and only if there does not exist a nonzero z-stable UV stable under all the maps xi-1. This is easily seen to be equivalent to the condition that ϕ-1(z) belongs to Ms(v,w)×T×G(v). ∎

Corollary A.7.

Given v,w let v,w be as in (A.5). Let θ be a generic character as in (A.7). Then we have an isomorphism ψ:M(v,w)Mθ(v,w) and an open immersion i:M(v,w)Mθ(v,w).∎

A.4.4 The compactification of local quiver varieties

From now on we fix u, y, u, y and θ as above. The existence of such a character θ is clear. For simplicity, we abbreviate

𝔐=𝔐(u,y),
𝔐=𝔐θ(u,y),
𝔐,=𝔐θ(u,y).

It is straightforward to check from (A.6) that the -actions on 𝔐 and 𝔐 given by

t(x¯,a¯)=(x,tx*,a,ta*),
t(x¯,a¯)=(x,tx*,,a,ta*,)

are compatible with the open embedding i in Corollary A.7. Applying (A.4) to the quivers Q and Q, we get the decompositions into smooth disjoint subvarieties

𝔐=v,w𝔐(v,w)#,
(𝔐)=v,w𝔐(v,w)#,

where v, w, v, w run over the sets of dimension vectors in I# and I,# such that

lvi,l=ui,Lwi,l=yi,ui=lvi,l,yi=lwi,l.

The variety 𝔐(v,w)# is realized in the same way as 𝔐(v,w)# by replacing the quiver Q by the quiver Q throughout. By taking the fixed points under the -actions, the map i yields an open immersion

(A.8)i:𝔐(v,w)#𝔐(v,w),#,

where v,w are defined by

vi1,l=vi2,l=vi,l,wi1,l=wi,l,wi2,l=0.

Since the quiver Q has no oriented cycle, each 𝔐,# is projective and we have obtained in this way the desired compactification of 𝔐#.

A.4.5 Proof of Theorem A.3

Now, let us consider the resolution of the diagonal. The smooth variety 𝔐×𝔐 carries T×G(w)×𝔾m-equivariant universal and tautological bundles given, for all iI, by

𝒱i,1=𝒱i𝒪𝔐,𝒱i,2=𝒪𝔐𝒱i,𝒲i,1,𝒲i,2.

Note that 𝒲i={0} for iI2 and that 𝒲i,1=𝒲i,2 is trivial as a vector bundle. We will simply denote this bundle by 𝒲i. We set also 𝒱=iI𝒱i and 𝒲=iI𝒲i. Consider the complex of equivariant bundles on 𝔐×𝔐 given by

(A.9)HomI(𝒱,1,𝒱,2)𝐚C𝐛HomI(𝒱,1,𝒱,2)[1],

where

  1. C=hΩ¯Hom(𝒱h,1,𝒱h′′,2)[εh]HomI(𝒱,1,𝒲)[1]HomI(𝒲,𝒱,2),

  2. εh=1 if hΩ,* and 0 otherwise,

  3. [1] denotes a grading shift with respect to the 𝔾m-action,

  4. the maps 𝐚,𝐛 are defined as

    𝐚((ξi))=(x,2ξh-ξh′′xh,1)(ai*,,2ξi)(-ξiai,1)

    and

    𝐛(yh,γi,δi)=hΩ(xh,2yh*-yh*xh,1)+i(ai,2δi+γia*,,1).

Lemma A.8.

Over the open subset M,×M the map a is injective and the map b is surjective.

Proof.

Fix tuples z1𝔐, and z2𝔐. Set z1=(x¯1,a¯1) and z2=(x¯2,a¯2). We will first prove that the map 𝐚 is injective over the point (z1,z2). Let (ξi)ker(𝐚) and put

Ki=ker(ξi)Vi,1,Ji=Im(ξi)Vi,2,
K=iKiW,J=iJi.

Since (ξi)ker(𝐚), the space K is a subrepresentation of z1 while I is a subrepresentation of z2. Since z1𝔐,, we have dim(Ki1)dim(Ki2) and hence dim(Ji1)dim(Ji2) for all iI. We deduce that

iIθidim(Ji)=iI[θi1(dim(Ji1)-dim(Ji2))+dim(Ji2)]0

with a strict inequality if there exists j such that Jj2{0} or dim(Jj1)>dim(Jj2), i.e., if J{0}. By θ-semistability of z2 we conclude that J={0} and that 𝐚 is injective.

To prove that 𝐛 is surjective, we dualize the previous argument. Using the trace pairing, we may view the transpose 𝐛* of 𝐛 as a map

𝐛*:HomI(𝒱,2,𝒱,1)[-1]C*,

where

C*=hΩ¯Hom(𝒱h′′,2,𝒱h1,1)[-εh]iIHom(𝒱i,2,𝒲i)[-1]iIHom(𝒲i,𝒱i,1).

Let (ξi)ker(𝐛*) and set

Ki=ker(ξi)Vi,2,Ji=Im(ξi)Vi,1,
K=iKiW,J=iJi.

Because (ξi)ker(𝐛*), the spaces J and K are respectively subrepresentations of z1 and z2. Since z1𝔐,, it follows that dim(Ji1)dim(Ji2), hence dim(Ki1)dim(Ki2) for all i. If dim(Kj1)>dim(Kj2) for some j, then we have

iIθidim(Ki)+θ=iI[θi1(dim(Ki1)-dim(Ki2))+dim(Ki2)]+θ
θj1-ivi>0,

which would contradict the semistability of z2. Thus dim(Ji1)=dim(Ji2) for all i. But then

iIθidim(Ji)=iIdim(Ji1)0

with strict inequality as soon as Ji{0} for some i. By the semistability of z1, this forces Ji={0} for all i, i.e., we have (ξi)=0 as wanted. Lemma A.8 is proved. ∎

By Lemma A.8 the restriction of the complex (A.9) to 𝔐,×𝔐 is quasi-isomorphic to an equivariant vector bundle =ker(𝐛)/Im(𝐚). Moreover, the bundle carries a section which vanishes precisely on the diagonal Δ𝔐, in

𝔐,×𝔐,𝔐,×𝔐.

Observe that Δ𝔐, is a closed subset of 𝔐,×𝔐. Indeed, see [28], the map

s=(0-a,2a*,,1)

defines a section of the vector bundle which vanishes precisely over the set of pairs (z1,z2) such that Hom(z1,z2){0} as kQ¯-modules. By the semistability condition, both z1 and z2 are simple so that Hom(z1,z2){0} if and only if z1=z2. Thus the zero set of s is Δ𝔐,.

As a consequence, we have the following equality in A*T×G(w)×𝔾m(𝔐,×𝔐):

[Δ𝔐,]=i0(-1)ich(Λi)[𝔐,×𝔐].

Let us now fix dimension vectors v, w for Q# and denote by v, w the associated dimension vectors for Q. We abbreviate 𝔐#=𝔐(v,w)# and 𝔐,#=𝔐(v,w)# as before. Let # be the restriction of to 𝔐#×𝔐,# under the map (A.8). We have

[Δ𝔐#]=i0(-1)ich(Λi#)[𝔐#×𝔐,#].

Consider the diagram

in which δ:𝔐#𝔐#×𝔐# is the diagonal embeddings and p1 is the projection onto the first factor. The three maps δ, δ and p1 are proper.

Let αA*T×G(w)(𝔐#) and choose αA*T×G(w)(𝔐,#) such that i*(α)=α. We have, on the one hand,

(A.10)p1*(([𝔐#]α)[Δ𝔐#])=p1*(([𝔐#]α)δ*[𝔐#])
=p1*δ*δ*([𝔐#]α)
=δ*([𝔐#]α)
=i*(α)
=α.

On the other hand, using the Giambelli formula we may write

i0(-1)ich(Λi#)=kakbk

for some classes ak belonging to the subring of AT×G(w)*(𝔐#) generated by the classes of the tautological bundles 𝒱i,l and 𝒲i,l as the pair (i,l) runs over I#, and some classes bk belonging to the subring of AT×G(w)*(𝔐,#) generated by the classes of the tautological bundles 𝒱i,l and 𝒲i,l as (lδi) runs over I,#. We have isomorphisms i*(𝒱ik,l)𝒱i,l and 𝒲i,l𝒲i,l. Further, we have

(A.11)p1*(([𝔐#]α)[Δ𝔐#])=kp1*(([𝔐#]α)(akbk))
=kakp*(αbk),

where p:𝔐,# is the map to a point.

Combining (A.10) and (A.11), we deduce that

Av,w#=A*T×G(w)(𝔐#).

The statement for the equivariant homology groups can be deduced from that for the Chow homology groups. Indeed, by the same argument as in Proposition 4.3 (c), since there is a contracting 𝔾m-action on 𝔐# whose fixed point subvariety is a projective graded quiver variety, the cycle map

cl:A*T×G(w)(𝔐#)H2*T×G(w)(𝔐#)

is surjective. This map preserves the tautological subrings. Hence, we have

Hv,w#=H*T×G(w)(𝔐(v,w)#).

Theorem A.3 is proved.

Index of notations

  1. H*, A*, M*, ε=1 or 2, H*G, A*G, M*G,

  2. f,

  3. (f,f)!, cl:A*G(X)H2*G(X), ch=l0chl,

  4. Q, Q*, Q¯, Ω, Ω*, Ω¯, ε(h), h, h′′, qij, q¯ij, qi, Ωij, Ω¯ij, ,, (,), dv, δi, kQ, x¯=(x,x*), R(v),

  5. G(v), 𝔤(v), μ:R(v)𝔤(v), M(v), Π, ,Π,

  6. θ(z), θ*(z), T=Tsp×Tdil,

  7. ΛW, Λν, Λ(v), =0 or 1,

  8. ΛW, Λν ,ΛF, Λν, Λ(v),

  9. J(i,q), μν,

  10. R(v,w), HomI(V,W), G(w), x¯, a¯, M(v,w), ρθ, Rθ(v,w), Mθ(v,w), Ms(v,w), dv,w, 𝔐0(v,w), 𝔐(v,w), πθ:𝔐θ(v,w)𝔐0(v,w), 𝔐(v,w)𝔸1, 𝔐0(v,w)𝔸1, v~, τ, κv,w, Q~=(I~,Ω~), RT(v,w), 𝔐0(τ), M(τ), Ms(τ), 𝔐(τ), 𝔏(v,w), M[ρ], L[ρ], 𝔐[ρ], 𝔏[ρ],

  11. H[v1,v2:w], 𝔥[v1,v2:w], H[v1,S:w], 𝔥[v1,S:w], H[v1,S], 𝔥[v1,S],

  12. Ls(v,w), 𝔏(v,w), :𝔾m×𝔐(v,w)𝔐(v,w), :𝔾m×𝔐(v,w)𝔐(v,w),

  13. H(w)G(w),

  14. 𝕜, 𝕜[v],

  15. H(v)G(v), 𝕜v, K,

  16. 𝕏, 𝕏+, Y, YA, YM, Y(v), Y(v,k),

  17. Y

  18. λ(q,z), τ=rank(T), Av(t),

  19. εi(x¯), R(v)l,i, Λ1(v)l,i, Y1(v)l,i, xi,l, xl,

  20. 𝐘, 𝕜(), 𝐘(0), 𝐘K, 𝒰i, Ir, Ie, Ih,

  21. Fw, Fw(v), Fw(v,k), Aw, Aw(v), Aw(v,k), (v1,v2:w), i,l(w), i,l, 𝔐(w), 𝒱i, ψi,l.

Acknowledgements

We would like to thank T. Bozec, F. Charles, M. McBreen, B. Davison, H. Nakajima, A. Negut and A. Okounkov for useful discussions and correspondences. Special thanks are due to B. Crawley-Boevey for providing us the proof of Proposition 3.1.

References

[1] A. Beilinson and V. Drinfeld, Quantization of Hitchin’s integrable system and Hecke eigensheaves, preprint, http://www.math.uchicago.edu/~mitya/langlands/hitchin/BD-hitchin.pdf. Search in Google Scholar

[2] T. Bozec, Quivers with loops and generalized crystals, Compos. Math. 152 (2016), no. 10, 1999–2040. 10.1112/S0010437X1600751XSearch in Google Scholar

[3] M. Brion, Poincaré duality and equivariant (co)homology, Michigan Math. J. 48 (2000), 77–92. 10.1307/mmj/1030132709Search in Google Scholar

[4] N. Chriss and V. Ginzburg, Representation theory and complex geometry, Mod. Birkhäuser Class., Birkhäuser, Boston 2010. 10.1007/978-0-8176-4938-8Search in Google Scholar

[5] D. H. Collingwood and W. M. McGovern, Nilpotent orbits in semisimple lie algebras, Van Nostrand Reinhold Math. Ser., Van Nostrand Reinhold, New York 1993. Search in Google Scholar

[6] W. Crawley-Boevey, On the exceptional fibres of Kleinian singularities, Amer. J. Math. 122 (2000), no. 5, 1027–1037. 10.1353/ajm.2000.0036Search in Google Scholar

[7] W. Crawley-Boevey, Geometry of the moment map for representations of quivers, Compos. Math. 126 (2001), no. 3, 257–293. 10.1023/A:1017558904030Search in Google Scholar

[8] W. Crawley-Boevey, Normality of Marsden–Weinstein reductions for representations of quivers, Math. Ann. 325 (2003), no. 1, 55–79. 10.1007/s00208-002-0367-8Search in Google Scholar

[9] B. Davison, The integrality conjecture and the cohomology of the preprojective stacks, preprint (2017), https://arxiv.org/abs/1602.02110v2. Search in Google Scholar

[10] B. Davison and S. Meinhardt, Cohomological Donaldson–Thomas theory of a quiver with potential and quantum enveloping algebras, preprint (2016), https://arxiv.org/abs/1601.02479. 10.1007/s00222-020-00961-ySearch in Google Scholar

[11] A. Douady and J.-L. Verdier, Séminaire de géométrie analytique, Astérisque 36–37, Société Mathématique de France, Paris 1976. Search in Google Scholar

[12] D. Edidin and W. Graham, Equivariant intersection theory, Invent. Math. 131 (1998), 595–634. 10.1007/s002220050214Search in Google Scholar

[13] D. Eisenbud and J. Harris, Intersection theory in algebraic geometry, Cambridge University Press, Cambridge 2016. Search in Google Scholar

[14] P. Etingof and C.-H. Eu, Koszulity and the Hilbert series of preprojective algebras, Math. Res. Lett. 14 (2007), no. 4, 589–596. 10.4310/MRL.2007.v14.n4.a4Search in Google Scholar

[15] W. Fulton, Intersection theory, 2nd ed., Ergeb. Math. Grenzgeb. (3), Springer, Berlin 1998. 10.1007/978-1-4612-1700-8Search in Google Scholar

[16] M. Goresky, R. Kottwitz and R. MacPherson, Equivariant cohomology, Koszul duality, and the localization theorem, Invent. Math. 131 (1998), no. 1, 25–83. 10.1007/s002220050197Search in Google Scholar

[17] W. Graham, Positivity in equivariant Schubert calculus, Duke Math. J. 109 (2001), no. 3, 599–614. 10.1215/S0012-7094-01-10935-6Search in Google Scholar

[18] T. Hausel, Kac’s conjecture from Nakajima quiver varieties, Invent. Math. 181 (2010), no. 1, 21–37. 10.1007/s00222-010-0241-3Search in Google Scholar

[19] T. Hausel and F. Rodriguez-Villegas, Mixed Hodge polynomials of character varieties, Invent. Math. 174 (2008), no. 3, 555–624. 10.1007/s00222-008-0142-xSearch in Google Scholar

[20] W. H. Hesselink, Polarizations in the classical groups, Math. Z. 160 (1978), no. 3, 217–234. 10.1007/BF01237035Search in Google Scholar

[21] M. Kontsevich and Y. Soibelman, Cohomological Hall algebra, exponential Hodge structures and motivic Donaldson–Thomas invariants, Commun. Number Theory Phys. 5 (2011), no. 2, 231–352. 10.4310/CNTP.2011.v5.n2.a1Search in Google Scholar

[22] M. Levine and F. Morel, Algebraic cobordism, Springer Monogr. Math., Springer, Berlin 2007. Search in Google Scholar

[23] A. Malkin, V. Ostrik and M. Vybornov, Quiver varieties and Lusztig’s algebra, Adv. Math. 203 (2006), no. 2, 514–536. 10.1016/j.aim.2005.05.002Search in Google Scholar

[24] D. Maulik and A. Okounkov, Quantum groups and quantum cohomology, preprint (2012), https://arxiv.org/abs/1211.1287. Search in Google Scholar

[25] K. McGerty and T. Nevins, Kirwan surjectivity for quiver varieties, Invent. Math. 212 (2018), no. 1, 161–187. 10.1007/s00222-017-0765-xSearch in Google Scholar

[26] S. Mozgovoy, Motivic Donaldson–Thomas invariants and McKay correspondence, preprint (2011), https://arxiv.org/abs/1107.6044. Search in Google Scholar

[27] S. Mozgovoy and O. Schiffmann, Counting Higgs bundles, preprint (2014), https://arxiv.org/abs/1411.2101. Search in Google Scholar

[28] H. Nakajima, Instantons on ALE spaces, quiver varieties, and Kac–Moody algebras, Duke Math. J. 76 (1994), no. 2, 365–416. 10.1215/S0012-7094-94-07613-8Search in Google Scholar

[29] H. Nakajima, Quiver varieties and Kac–Moody algebras, Duke Math. J. 91 (1998), no. 3, 515–560. 10.1215/S0012-7094-98-09120-7Search in Google Scholar

[30] H. Nakajima, Quiver varieties and finite-dimensional representations of quantum affine algebras, J. Amer. Math. Soc. 14 (2001), no. 1, 145–238. 10.1090/S0894-0347-00-00353-2Search in Google Scholar

[31] H. Nakajima, Quiver varieties and t-analogs of q-characters of quantum affine algebras, Ann. of Math. (2) 160 (2004), no. 3, 1057–1097. 10.4007/annals.2004.160.1057Search in Google Scholar

[32] H. Nakajima, Quiver varieties and branching, SIGMA Symmetry Integrability Geom. Methods Appl. 5 (2009), Paper No. 003. 10.3842/SIGMA.2009.003Search in Google Scholar

[33] A. Neguţ, Exts and the AGT relations, Lett. Math. Phys. 106 (2016), no. 9, 1265–1316. 10.1007/s11005-016-0865-3Search in Google Scholar

[34] A. Okounkov, On some interesting Lie algebras, Talk at the conference in honor of V. Kac, IMPA (2013), http://www.math.columbia.edu/~okounkov/. Search in Google Scholar

[35] C. A. M. Peters and J. H. M. Steenbrink, Mixed Hodge structures, Ergeb. Math. Grenzgeb. (3) 52, Springer, Berlin 2008. Search in Google Scholar

[36] J. Ren and Y. Soibelman, Cohomological Hall algebras, semicanonical bases and Donaldson–Thomas invariants for 2-dimensional Calabi–Yau categories (with an appendix by Ben Davison), Algebra, geometry, and physics in the 21st century, Progr. Math. 324, Birkhäuser, Cham (2017), 261–293. 10.1007/978-3-319-59939-7_7Search in Google Scholar

[37] O. Schiffmann, Variétés carquois de Nakajima, Séminaire Bourbaki. Volume 2006/2007. Exposés 967–981, Astérisque 317, Société Mathématique de France, Paris (2008), 295–344, Exposé No. 976. Search in Google Scholar

[38] O. Schiffmann and E. Vasserot, Hall algebras of curves, commuting varieties and Langlands duality, Math. Ann. 353 (2012), no. 4, 1399–1451. 10.1007/s00208-011-0720-xSearch in Google Scholar

[39] O. Schiffmann and E. Vasserot, Cherednik algebras, W-algebras and the equivariant cohomology of the moduli space of instantons on 𝐀2, Publ. Math. Inst. Hautes Études Sci. 118 (2013), 213–342. 10.1007/s10240-013-0052-3Search in Google Scholar

[40] O. Schiffmann and E. Vasserot, The elliptic Hall algebra and the K-theory of the Hilbert scheme of 𝔸2, Duke Math. J. 162 (2013), no. 2, 279–366. 10.1215/00127094-1961849Search in Google Scholar

[41] O. Schiffmann and E. Vasserot, On cohomological Hall algebras of quivers: Yangians, preprint (2017). 10.1515/crelle-2018-0004Search in Google Scholar

[42] M. Varagnolo, Quiver varieties and Yangians, Lett. Math. Phys. 53 (2000), no. 4, 273–283. 10.1023/A:1007674020905Search in Google Scholar

[43] Y. Yang and G. Zhao, Cohomological Hall algebra of a preprojective algebra, preprint (2015), https://arxiv.org/abs/1407.7994v2. 10.1112/plms.12111Search in Google Scholar

Received: 2017-06-24
Revised: 2018-01-22
Published Online: 2018-05-03
Published in Print: 2020-03-01

© 2021 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 27.4.2024 from https://www.degruyter.com/document/doi/10.1515/crelle-2018-0004/html
Scroll to top button