152
Views
0
CrossRef citations to date
0
Altmetric
Research Article

Moduli of representations of one-point extensions

&
Received 05 Nov 2022, Accepted 30 Apr 2024, Published online: 19 May 2024

Abstract

We study moduli spaces of (semi-)stable representations of one-point extensions of quivers by rigid representations. This class of moduli spaces unifies Grassmannians of subrepresentations of rigid representations and moduli spaces of representations of generalized Kronecker quivers. With homological methods, we find numerical criteria for non-emptiness and results on basic geometric properties, construct generating semi-invariants, expand the Gel’fand MacPherson correspondence, and derive a formula for the Poincaré polynomial in singular cohomology of these moduli spaces.

2020 Mathematics Subject Classification:

1 Introduction

In this paper we construct and study moduli spaces parameterizing isomorphism classes of representations of so-called one-point extensions of path algebras of quivers. This constitutes a class of algebras of global dimension two, for which many of the favorable properties of moduli spaces of representations of quivers still hold. Namely, we find numerical criteria for non-emptiness and results on basic geometric properties, construct generating semi-invariants, expand the Gel’fand MacPherson correspondence, and derive a formula for the Poincaré polynomial in singular cohomology of these moduli spaces. We explicitly apply the developed theory in several examples.

To do this, we fix a path algebra A = kQ of a finite quiver, which we extend by a representation T of A to the one-point extension algebra A[T]. We construct standard projective resolutions for representations of A[T]. One of the most important consequences of this is an explicit description of the space Ext2 of representations, which allows us to conclude its vanishing on so-called full representations (under the assumption of T being rigid). See Theorems 3.1, 3.3, 3.4, and 3.6 for precise formulations. Moreover, the standard resolutions allow us to calculate the Euler form of A[T] in Theorem 3.7, Corollaries 3.8 and 3.10. After these preparations, we consider the representation varieties of A[T], interpret the found homological properties in this geometric setting, and rediscover some results by Schofield and Crawley-Boevey (with different methods) in Theorem 4.3, Corollaries 4.4 and 4.5. Moreover, in this way we can determine the Zariski tangent space of the representation variety in each point (Theorem 4.1) and conclude that the open subset of full representations is smooth and irreducible (Theorem 4.8).

We follow the GIT approach of King in the construction of moduli spaces. For this, we choose a canonical stability condition, such that the resulting spaces unify quiver Grassmannians of subrepresentations of rigid representations (Theorem 7.1) and moduli spaces of representations of generalized Kronecker quivers. We find a numerical criterion for semistability in Theorem 5.2, which allows us to conclude that semi-stable representations are full representations (Corollary 5.7). In this way, we can apply the above geometric properties of representation varieties to prove that the resulting moduli spaces are smooth, irreducible and of expected dimension in Theorem 5.8.

After this, we prove a relative version of a recursive numerical criterion [Citation9] for non-emptiness of the semi-stable locus in Theorems 5.9 and 5.10. A set of generators for the ring of semi-invariants, closely following Schofield and Van den Bergh ([Citation13]) is given in section 6.

Moreover, we find a form of Gel’fand MacPherson correspondence in terms of these moduli spaces in Theorem 7.1. We finish the paper by deriving a recursive formula to determine the Poincaré polynomial of the moduli spaces (Theorem 8.2).

2 One-point extensions and their representations

We fix an algebraically closed field k of characteristic zero. Let A = kQ be the path algebra of a finite quiver Q, and let T be a finite dimensional A-module. We consider the one-point extension of A by T A[T]:=(AT0k):={(at0λ) : aA,tT,λk}.

Recall that the multiplication in (the k-algebra) A[T] is given by the formal matrix multiplication (at0λ)·(at0λ):=(aaat+tλ0λλ) for a,aA,λ,λk,t,tT and componentwise addition.

We define a category RepA(Tk?2?1) as follows. Its objects M are tuples (M1, M2), where M1 is a left A-module and M2 is a k-vector space, together with a map of A-modules fM:TkM2M1. A morphism φ:MN is a tuple (φ1,φ2), where φ1:M1N1 is a map of A-modules and φ2:M2N2a k-linear map, such that the diagram

commutes. Composition of morphisms is defined componentwise.

Lemma 2.1.

The category RepA(Tk?2?1) is equivalent to the category of left A[T]-modules.

Proof.

See for example [Citation11]. □

We can easily describe a quiver Q̂ and relations R in this situation such that A[T]kQ̂/R. By extending Q the following way Q̂0=Q0{},Q̂1=Q1 {ρl,(i):i : iQ0,l=1,,dimTi}, we obtain the quiver Q̂. We obtain the relations R by using the transformation matrices induced by the fixed representation T: For each vertex iQ0 we choose a k-basis ρ1,(i),,ρdimTi,(i) of Ti and write T(α)ρl,(i)=s=1dimTjλs,l(α)ρs,(j)

for all l=1,,dimTi and each arrow (α:ij)Q1. The coefficients of these linear combinations induce the relations R (1) α·ρl,(i)=s=1dimTjλs,l(α)ρs,(j).(1)

Obviously this construction of relations does not depend (up to isomorphism of k-algebras) on the basis we choose in Ti for each vertex iQ0.

Example 2.1.

Let Q=(12).

By extending the path algebra of Q with k3[1 0 0]k we get

3 Homological properties

Since A is the path algebra of a quiver Q, we can and will identify A with the tensor algebra TRX, where R is the semisimple k-algebra generated by the vertices of Q and X is the R-R-bimodule generated (as a k-vector space) by the arrows of Q.

3.1 The standard resolution

Let {e1,,en} be the complete set of primitive orthogonal idempotents given by the length 0 paths in A. Then (e1000),,(en000),(0001) is a complete set of primitive orthogonal idempotents of A[T]. Let R˜ be the k-subalgebra of A[T] given by R˜:=i=1nk(ei000)    k(0001).

The multiplication of A resp. A[T] induces the Eilenberg sequence [Citation2, Proposition 2.7.3] E(A):0Ω(A)κAARAμAA0, resp. E(A[T]):0Ω(A[T])A[T]R˜A[T]μA[T]A[T]0, with Ω(A):=ker(μA) resp. Ω(A[T]):=ker(μA[T]).

This sequence of A-bimodules resp. A[T]-bimodules splits in the category of right A-modules resp. A[T]-modules.

Theorem 3.1.

Let M˜=(M,V,f:TkVM) be an A[T]-module. Then there is a short exact sequence 0Ω(A[T])A[T]M˜A[T]R˜M˜M˜0, where A[T]R˜M˜=(TkV,V, TkVidTkV) (ARM,0,0),Ω(A[T])A[T]M˜=(TkV,0,0)(Ω(A)AM,0,0).

Proof.

We obtain the short exact sequence by tensoring the split short exact sequence of right A[T]-modules E(A[T]) over A[T] with M˜.

The first equation follows immediately using the definition of R˜. Namely, for M0:=V we have A[T]e˜0=(0T0k) and A[T]R˜M˜=i=0nA[T]e˜ikMi.

The second equation follows from the following commutative diagram with exact rows:

where E(A)AM:0Ω(A)AMκMARMμMM0, and g is defined as g((tv)+(ωm)):=tv1f(tv)+κM(ωm) for tvTkV,  ωmΩ(A)AM. □

Remark 3.2.

To simplify the notation, set ΩA=Ω(A) and N=ΩAAM. Combining Theorem 3.1 with the standard resolution 0ΩAAXAXX0 of A-modules, we obtain the following long exact sequence:

Using this construction, we obtain:

Theorem 3.3.

For A[T]-modules M˜=(M,V,f:TkVM) we have the standard projective resolution:

Here g and h are defined as g(a(tv)+axm)=atvaf(tv)+axmaxm,h(ax(tv))=(ax(tv)a(xtv))+axf(tv), for a(tv)AR(TkV),axm(ARXRA)AM and ax(tv)(ARXRA)A(TkV).

In particular, we have gldim A[T]2.

3.2 Characterizing Ext2

Let M˜=(M,V,f:TkVM),N˜=(N,W,g:TkWN) be A[T]-modules.

By using the Eilenberg sequence we obtain the following commutative diagram with exact rows and columns

Note that, since R is a semisimple k-algebra, Ω(A) and ΩAAL are projective modules for all A-modules L.

Applying HomA(?,N), we obtain the following commutative diagram

Now we consider the standard projective resolution of M˜ of Theorem 3.3 P(M˜)M˜0 and the induced cochain complex C:=HomB(P(M˜),N˜): C:0HomR˜(M˜,N˜)HomA(ΩAAM,N)HomR(TkV,N)h*HomA(ΩAA(TkV),N)0

Then Hi(C)=ExtA[T]i(M˜,N˜) for i = 0, 1, 2.

We have h*=HomA[T](h,N˜)=HomA[T]([κTkV   ΩAf],N˜)=[κTV*   (ΩAf)*].

Moreover, we have the following commutative diagrams:

Thus we obtain: (2) im((ΩAf)*)=im((ΩAf)*).(2)

We sum up and obtain finally: ExtA[T]2(M˜,N˜)=HomA(ΩAA(TkV),N)/im(h*)=HomA(ΩAA(TkV),N)/im([κTV*   (ΩAf)*])=(2)(HomA(ΩAA(TkV),N)/im((ΩAf)*))/im(κTV*¯)=HomA(ΩAAker(f),N)/im(κTV*¯)=ExtA(ker(f),N).

We thus arrive at the following description of ExtA[T]2:

Theorem 3.4.

For A[T]-modules M˜=(M,V,f:TkVM), N˜=(N,W,g:TkWN) there is an isomorphism: ExtA[T]2(M˜,N˜)˜ ExtA(kerf,N).

In particular, if T is projective, so is ker(f), thus ExtA[T]2 vanishes identically. In other words, A[T] is again hereditary in this case.

Definition 3.5.

We call an A[T]-module M˜=(M,V,f:TkVM) full, if f is a surjective map.

Theorem 3.6.

Assume ExtA(T,T)=0. Then, for full A[T]-modules M˜,N˜, the vanishing property ExtA[T]2(M˜,N˜)=0 holds.

Proof.

Write M˜=(M,V,f:TkVM) and N˜=(N,W,g:TkWN). We consider 0ker(g)TkWgN0 and apply HomA(ker(f),?) and obtain the long exact sequence 0HomA(ker(f),ker(g))HomA(ker(f),TkW)HomA(ker(f),N)ExtA(ker(f),ker(g))ExtA(ker(f),TkW)ExtA(ker(f),N)0.

Now we will show that ExtA(ker(f),N)=0 holds by showing that ExtA(ker(f),TkW)=0 holds. We also have: 0ker(f)TkVfM0.

Applying HomA(?,TkW) to this, we obtain the long exact sequence: 0HomA(M,TkW)HomA(TkV,TkW)HomA(ker(f),TkW)ExtA(M,TkW)ExtA(TkV,TkW)ExtA(ker(f),TkW)0.

Since ExtA(T,T)=0, in particular ExtA(TkV,TkW)=0 holds.

So ExtA(ker(f),TkW)=0 and therefore ExtA(ker(f),N)=0. Using Theorem 3.4, we conclude the proof. □

3.3 Derivations and the Euler form of A[T]

Let M˜=(M,V,f:TkVM),N˜=(N,W,g:TkWN) be A[T]-modules. To shorten notation, we define B=A[T]. In this section we determine the dimension of a space of derivations dimDerR˜(B,HomR˜(N˜,M˜)).

To do this, we consider the canonical exact sequence 0HomB(N˜,M˜)HomR˜(N˜,M˜)cDerR˜(B,HomR˜(N˜,M˜))ExtB(N˜,M˜)0, where c is defined as (βi:N˜iM˜i)iQ0α:ijαQ1M˜αβiβjN˜α.

We obtain the equality: dimDerR˜(B,HomR˜(N˜,M˜))=(dimHomB(N˜,M˜)dimExtB(N˜,M˜))+HomR˜(N˜,M˜).

On the other hand, we have the following description:

We consider the standard projective resolution of N˜ (Theorem 3.3) P(N˜)N˜0 and the induced cochain complex C:=HomB(P(N˜),M˜): C:0HomR˜(N˜,M˜)HomA(ΩAAN,M)HomR(TkW,M)h*HomA(ΩAA(TkW),M)0

Then Hi(C)=ExtA[T]i(N˜,M˜)(i=0,1,2).

This way we obtain the equality:

dimHomB(N˜,M˜)dimExtB(N˜,M˜)+dimExtB2(N˜,M˜)=dimHomR˜(N˜,M˜)dimHomA(ΩAAN,M)dimHomR(TkW,M)+dimHomA(ΩAA(TkW),M). We easily determine: dimHomA(ΩAAN,M)=α:ijαQ1dimNi·dimMj,dimHomR(TkW,M)=dimW·iQ0dimTi·dimMi,dimHomA(ΩAA(TkW),M)=dimW·α:ijαQ1dimTi·dimMj.

By using the characterization of ExtA[T]2 of Theorem 3.4, we end up in:

Theorem 3.7.

For finite-dimensional A[T]-modules M˜=(M,V,f:TkVM) and N˜=(N,W,g:TkWN) we have dimDerR˜(B,HomR˜(N˜,M˜))=dimDerR(A,HomR(N,M))+dimW·T,MA+dimExtA(ker(g),M), where .,.A denotes the homological Euler-form of A.

Corollary 3.8.

If we assume ExtA(T,T)=0, then for full A[T]-modules M˜ and N˜ we have dimDerR˜(B,HomR˜(N˜,M˜))=dimDerR(A,HomR(N,M))+dimW·T,MA.

Proof.

The claim follows immediately using Theorems 3.7 and 3.6. □

We notate a dimension vector dNQ̂0 as a tuple (s, d), where s=d and dNQ0.

Definition 3.9.

For dimension vectors (s,d),(s,d)NQ̂0 we set (s,d),(s,d)A[T]:=sss·dim¯T,dQ+d,dQ.

Corollary 3.10.

For the homological Euler-form of A[T] the following identity holds: M˜,N˜A[T]=dim¯M˜,dim¯N˜A[T], where M˜,N˜ are (finite-dimensional) A[T]-modules and .,.Q denotes the Euler-form of Q.

Proof.

The identity follows from the above discussion. □

4 Varieties of representations of one-point extensions

For all standard notions on varieties of representations of algebras, we refer to [Citation5]. Let (s, d) be a dimension vector of Q̂. Using the isomorphism A[T]kQ̂/(R), we can realize the variety of representations of A[T] with dimension (s, d) as a (Zariski-)closed subvariety of the variety of representations of Q̂ with dimension (s, d) denoted by Rep(s,d)(Q̂): Rep(s,d)(A[T])clsdRep(s,d)(Q̂).

4.1 The Zariski-tangent space

For M˜Rep(s,d)(A[T]) we have (see [Citation5, Example 3.10.]): TM˜Rep(s,d)(A[T])=DerR˜(A[T],EndR˜(M˜)).

We set t:=dim¯ T and get by using Theorem 3.7:

Theorem 4.1.

For M˜=(M,V,f:TkVM)Rep(s,d)(A[T]) we have dimTM˜Rep(s,d)(A[T])=dimRepd(Q)+s·t,dQ+dimExtA(ker(f),M).

4.2 A well-behaved subvariety in the rigid case

We consider Rep(s,d)full(A[T]):={(f,M) : MRepd(Q),  rank(f)=γTs,d}open Rep(s,d)(A[T]), where γTs,dNQ0 denotes the general rank of homomorphism from T s to a representation of dimension d (see [Citation12, Section 5]).

Theorem 4.2.

If we assume γTs,d=d and ExtA(T,T)=0, then Rep(s,d)full(A[T]) is smooth and every component has dimension dimRepd(Q)+s·dim¯ T,dQ, i.e. Rep(s,d)full(A[T]) is a local complete intersection.

Proof.

By counting the explicit defining polynomial equations Rep(s,d)(A[T])clsdRep(s,d)(Q̂), induced by (1), we find that every (nonempty) component of Rep(s,d)(A[T]) has dimension dim Repd(Q)+s·dim¯ T,dQ.

On the other hand, for each M˜Rep(s,d)full(A[T]) we have dimM˜Rep(s,d)full(A[T])dimTM˜Rep(s,d)(A[T]).

The claim follows immediately by using the dimension formula in Theorem 4.1. □

4.3 On homomorphisms from a fixed representation

We denote the open dense subset {MRepd(Q) : dimHomA(T,M)=hom(T,d)and γT,M=γT,d} of Repd(Q) by VT,d. Here, hom(T,d) is the dimension of the space of homomorphisms from the fixed representation T to a general representation of dimension vector d (see [Citation3]). Moreover, γT,M is the unique maximal rank of homomorphisms from T to M.

We consider the regular map π˜:Rep(1,d)full(A[T])Repd(Q),(f,M)M,

which induces a regular map π:VVT,d, where V is the open preimage π˜1(VT,d)Rep(1,d)full(A[T]).

Obviously VT,d is irreducible and the fibers of π are irreducible of dimension hom(T,d). So there is a unique irreducible component V0 of V of maximal dimension which dominates π, that is, we have π(V0)¯=VT,d (Proposition 4.7)and dimV=dimV0=dimRepd(Q)+hom(T,d).

Since the regular points in V form an open dense subset, there is regular point M˜ in the irreducible component V0, and we have: dimV0=dimM˜V=dimTM˜V=dimTM˜Rep(1,d)full(A[T]).

Using the description of the tangent space by derivations, we thus find:

Theorem 4.3.

We have hom(T,d)=t,dQ+dimExtA(ker(f),M) for a general A-module homomorphism f:TM.

As a side remark, we rediscover the following result of Schofield [Citation12] and Crawley-Boevey [Citation3]:

Corollary 4.4.

We have hom(T,d)=γT,d,dQ+dimHomA(ker(f),M) for a general A-module homomorphism f:TM.

Proof.

Use the Euler form of Q and the identity dimHomA(ker(f),M)dimExtA(ker(f),M)=dim¯ ker(f),d.

Corollary 4.5.

If we assume γT,d=d and ExtA(T,T)=0, then hom(T,d)=t,dQ.

4.4 An irreducible component in the rigid case

We need the following facts from algebraic geometry:

Proposition 4.6.

Let π:XY be a regular map of affine varieties. Then there is an open dense subset UX such that for all xU Txπ1(π(x))=ker(dπx).

Proof.

See for example [Citation4]. □

Proposition 4.7.

Let f:XY be a regular map of quasi-projective varieties.

If Y is irreducible and all fibers of f are irreducible and of same dimension d (in particular f is surjective), then:

  1. There is a unique irreducible component X0 of X that dominates Y, i.e. f(X0)¯=Y.

  2. Each irreducible component Xi of X is a union of fibers of f. Its dimension is equal to dimf(Xi)¯+d.

In particular, we can conclude X is irreducible if either of the following holds:

  1. X is equidimensional.

  2. f is closed.

Proof.

See [Citation6]. □

Theorem 4.8.

If we assume γTs,d=d and ExtA(T,T)=0, then Rep(s,d)full(A[T]) is irreducible and smooth of dimension dimRep(s,d)full(A[T])=dimRepd(Q)+s·t,dQ.

Proof.

By Theorem 4.2, it remains to prove that Rep(s,d)full(A[T]) is irreducible. We consider the dominant map π˜:Rep(s,d)full(A[T])Repd(Q),M˜=(M,ks,f)M.

For each M˜=(M,ks,f)Rep(s,d)full(A[T]) the regular map Repd(Q)Rep(s,d)(A[T]),N(N,ks,0) induces a split mono for the differential π˜ in M˜ dπ˜M˜:TM˜Rep(s,d)full(A[T])TMRepd(Q),

i.e. dπ˜M˜ is surjective. For each M˜=(M,ks,f)Rep(s,d)full(A[T]) we have HomA(Ts,M)TM˜π˜1(π˜(M˜)).

There is an open dense subset URep(s,d)full(A[T]) such that for each M˜U we have HomA(Ts,M)ker(dπ˜M˜)

(Theorem 4.6). Since dπ˜M˜ is surjective, by using Theorem 4.5 we obtain: dimHomA(Ts,M)=hom(Ts,d) for each M˜U. So U{(M,ks,f)Rep(s,d)full(A[T]) : hom(Ts,M)=dimHomA(Ts,M)}=:V.

Since URep(s,d)full(A[T]) is dense, VRep(s,d)full(A[T]) is dense, too.

The map π˜ induces a surjective regular map π:VVTs,d,(M,ks,f)M, where VTs,dRepd(Q) is dense (and thus irreducible). All fibers of π are irreducible of equal dimension and V is equidimensional, thus V has to be irreducible by Theorem 4.7. Using V¯=Rep(s,d)full(A[T]), we can conclude the claim. □

5 Semistability

For all notions concerning stability and moduli spaces of representations we refer to [Citation9]. For an A[T]-module M˜=(M,V,f:TkVM) we define its slope μ(M˜)=μ(M,V,f):=dimVdimV+dimM.

Definition 5.1.

An A[T]-module M˜=(M,V,f:TkVM) is called semi-stable (resp. stable) if μ(U)μ(M˜) (resp. μ(U)<μ(M˜)) for all proper subrepresentation 0UM˜.

5.1 A first criterion

Theorem 5.2.

Let M˜=(M,V,f:TkVM) be an A[T]-module with M˜0. Then M˜ is (semi-)stable iff for all subspaces WM˜ the following inequality is fulfilled: WM˜

Proof.

For all subobjects U˜=(U,W,g)M˜, we have: μ(U˜)<(=)μ(M˜)dimM·dimWdimV<(=)dimU.

We can easily determine the total space of the smallest subobject W¯ of M˜ containing a given WV. Namely, we have W¯i=l=1tiM˜ρl,(i)(W) for each iQ0. □

5.2 An observation on the Harder-Narasimhan filtration

At first we recall the notion of Harder-Narasimhan filtration.

Definition 5.3.

  1. A dimension vector (s,d)N×NQ0 is called (semi-)stable if there is a (semi-)stable representation of A[T] with dimension vector (s, d).

  2. A tuple

    (s,d)*=((s1,d1),,(sr,dr))(N×NQ0)r

    of dimension vectors is of HN-type if each (sl,dl) (l=1,,r) is semi-stable and

    μ(s1,d1)>>μ(sr,dr).

  3. A filtration

    0=M˜0M˜1M˜r=M˜

    of a representation M˜ (of A[T]) is called Harder-Narasimhan (HN) if each quotient M˜l/M˜l1 (l=1,,r) is semi-stable and

    μ(M˜1/M˜0)>μ(M˜2/M˜1)>>μ(M˜r/M˜r1).

Proposition 5.4.

Every representation M˜ of A[T] admits a unique Harder-Narasimhan filtration.

Proof.

See [Citation9]. □

Definition 5.5.

For a HN type (s,d)* we denote by Rep(s,d)*HN(A[T])Rep(s,d)(A[T]) the subset of representations whose HN filtration is of type (s,d)*. Rep(s,d)*HN(A[T]) is called HN stratrum for the HN type (s,d)*. More generally, we denote by Rep(s,d)(s,d)*(A[T])Rep(s,d)(A[T]) the subset of representations M˜ possessing a filtration of type (s,d)*, i.e. there is a chain of subrepresentations 0=M˜0M˜1M˜r=M˜ with dim¯ M˜l/M˜l1=(sl,dl) for l=1,,r.

Theorem 5.6.

Let M˜=(M,V,f:TkVM) be a representation of A[T] with V0. Then the following are equivalent:

  1. f:TkVM is surjective.

  2. For the HN filtration of M˜

    0=M˜0M˜1M˜r=M˜

    we have

    μ(M˜r/M˜r1)0.

Proof.

a) b): We denote U˜=M˜r1, which we write as U˜=(U,W,g:TkWU).

Since U˜M˜ is a subrepresentation, we have the following commutative diagram:

Assume μ(M˜/U˜)=0, i.e. dimVdimW=0. So we obtain W = V, and since f is surjective we can conclude from the above commutative diagram that U already equals M. In other words, U˜=M˜, contradicting our assumption. b) a): Assume the structure morphism f:TkVM is not surjective. Then we can consider the proper subobject U˜M˜ induced by the commutative diagram

Obviously then we have μ(M˜/U˜)=0, and since f is not surjective, we neither have dim¯ M˜/U˜=0. So M˜/U˜ is a semi-stable representation.

Now look at the HN filtration of U˜. Since the structure morphism of U˜ is surjective we can conclude from the first part of this proof that for the HN filtration of U˜=YlYl1Y1Y0=0 we have μ(Yl/Yl1)0.

Since by definition the slope is always 0 we can conclude μ(Yl/Yl1)>μ(M˜/U˜).

Using the uniqueness of the HN filtration we finally obtain that 0=Y0Y1Yl=U˜M˜ must be the HN filtration of M˜. But this contradicts our assumption about the HN filtration of M˜. So f has to be surjective. □

Consequences of Theorem 5.6 are:

Corollary 5.7.

For (s,d)N×NQ0 we have the following connection between the semi-stable representations and full representations: Rep(s,d)sst(A[T])Rep(s,d)full(A[T]).

5.3 Geometric consequences for the moduli space

The linear algebraic group G(s,d):=GLs(k)×iQ0GLdi(k)

acts on Rep(s,d)(Q̂) via the base change action (gi)i·(M˜α)αQ̂1=(gjM˜αgi1)α:ij.

Rep(s,d)(A[T]) is stable under this G(s,d)-action. By definition, the G(s,d)-orbits in Rep(s,d)(A[T]) correspond bijectively to isomorphism classes [M˜] of representations of A[T] of dimension vector (s, d). We consider the stability function Θ˜ for Q̂ given by Θ˜=1 and Θ˜i=0 for iQ0. The associated slope function on representations of Q̂ coincides with the slope function μ on A[T]-modules. This allows us to define moduli spaces M(s,d)sst(A[T]) resp. M(s,d)st(A[T]) as the algebraic quotient of Rep(s,d)sst(A[T]), resp. the geometric quotient of Rep(s,d)st(A[T]), by G(s,d).

Theorem 5.8.

Assume ExtA(T,T)=0 and γTs,d=d.

If Rep(s,d)st(A[T]), then both M(s,d)sst(A[T]) and M(s,d)st(A[T]) are irreducible and smooth of dimension dim M(s,d)sst(A[T])=1(s,d),(s,d)A[T].

Proof.

We calculate fiber dimensions for the geometric quotient π:Rep(s,d)st(A[T])M(s,d)st(A[T]), and use Corollary 5.7, Theorem 3.6 and the fact that the endomorphism rings of stable representations are trivial. □

Next, we introduce the Harder-Narasimhan stratification. Note that the term stratification is used in a weak sense, meaning a finite decomposition of a variety into locally closed subsets.

5.4 Harder-Narasimhan stratification

In this section we write Rep(s,d) for Rep(s,d)full(A[T]) to simplify notation.

Theorem 5.9.

Let assume ExtA(T,T)=0 and γTs,d=d.

The HN-strata for the HN-types (s,d)*=((s1,d1),,(sr,dr)) with weight (s, d), i.e. (s,d)=i=1r(si,di), and sr0 define a stratification of Rep(s,d).

The codimension of Rep(s,d)*HN in Rep(s,d) is given by: k<l(sk,dk),(sl,dl)A[T].

Proof.

Let F*:0=F0F1Fr be a flag of type (s,d)* in the Q̂0-graded vector space kd×ks=iQ0kdi×ks, i.e. Fl/Fl1kdl×ksl for l=1,,r, and denote by Fil the i-component of F l .

Denote by Z˜(s,d)* the closed subvariety of Rep(s,d)(A[T]) of representations M˜ which are compatible with F*, i.e. M˜γ(Fil)Fjl for l=1,,r and for all arrows (γ:ij) in Q̂. We have the regular map p˜(s,d)*:Z˜(s,d)*Rep(s1,d1)(A[T])××Rep(sr,dr)(A[T]) given by the projection p˜(s,d)* mapping M˜Z˜(s,d)* to the sequence of subquotients with respect to F*. The map p˜(s,d)* induces a regular map p(s,d)*:Z(s,d)*Rep(s1,d1)××Rep(sr,dr), where Z(s,d)*:=p˜(s,d)*1(Rep(s1,d1)××Rep(sr,dr))Z˜(s,d)* open.

A minute reflection shows that Z(s,d)*Rep(s,d), and it is a locally closed subset. By Theorem 4.8, Rep(s,d),Rep(s1,d1),,Rep(sr,dr) are irreducible.

We set e2*:=((s1,d1),(s2,d2)).

We obtain the regular map pe2*:Ze2*Rep(s1,d1)×Rep(s2,d2), with fibers at (M˜1,M˜2)Rep(s1,d1)×Rep(s2,d2) given by pe2*1(M˜1,M˜2)=DerR˜(B,HomR˜(M˜2,M˜1)).

Thus all fibers are irreducible and of equal dimension (Corollaries 5.7 and 3.8). By Theorem 4.7 we have dimZe2*=dimpe2*1(M˜1,M˜2)+dim(Rep(s1,d1)×Rep(s2,d2)).

Now we set e¯3*:=((s1,d1)+(s2,d2),(s3,d3)),e3*:=((s1,d1),(s2,d2),(s3,d3)).

Therefore we obtain the regular map pe¯3*:Ze¯3*Rep(s1,d1)+(s2,d2)×Rep(s3,d3).

Since Ze3*=pe¯3*1(Ze2*×Rep(s3,d3)),

the map pe¯3* induces a regular map Ze3*Ze2*×Rep(s3,d3).

As above, we see that this regular map has irreducible fibers of equal dimension, that is, for (M˜1,M˜3)Ze2*×Rep(s3,d3), we have: dimZe3*=dimDerR˜(B,HomR˜(M˜3,M˜1))+dimZe2*+dimRep(s3,d3).

Inductively, we obtain in this way a regular map Zer*Zer1*×Rep(sr,dr) with irreducible fibers of equal dimension, where er1*=((s1,d1),,(sr1,dr1)),er*=(s,d)*.

Summing up, for (M˜1,M˜r)Zer1*×Rep(sr,dr) this yields: dimZ(s,d)*=dimDerR˜(B,HomR˜(M˜r,M˜1))+dimZer1*+dimRep(sr,dr).

Together with the formula in Corollary 3.8, we find: dimZ(s,d)*=n<lα:ijαQ1dildjn+slt,dnQ+i=1rdim Rep(si,di).

To simplify the notation in the following, we write Z (resp. p) for Z(s,d)* (resp. p(s,d)*). The preimage of Rep(s1,d1)sst××Rep(sr,dr)sst

under p gives us an open subvariety Z0 of Z and Z˜. Since the varieties Rep(s1,d1),,Rep(sr,dr) are irreducible, we see in a similar manner to Z that dimZ0=dimZ holds.

The action of G(s,d) on Rep(s,d)(A[T]) induces actions of the parabolic subgroup P(s,d)* of G(s,d), consisting of elements fixing the flag F*, on Z0 and Z. The image of the associated fiber bundle G(s,d)×P(s,d)*Z˜ under the action morphism m equals Rep(s,d)(s,d)*(A[T]), which is thus a closed subvariety of Rep(s,d)(A[T]). The image of G(s,d)×P(s,d)*Z0 under m equals Rep(s,d)*HN, and G(s,d)×P(s,d)*Z0 is the full preimage. By the uniqueness of the HN filtration, the morphism m is bijective over Rep(s,d)*HN, which therefore is a locally closed subvariety of Rep(s,d)(A[T]).

The canonical map G(s,d)×P(s,d)*Z0G(s,d)/P(s,d)* is (Zariski) locally trivial. Therefore dimG(s,d)×P(s,d)*Z0=(dimG(s,d)P(s,d)*)+dimZ0.

The codimension of Rep(s,d)*HN in Rep(s,d) is now easily computed as n<l(sn,dn),(sl,dl)A[T], using the identity d,dQ=dimGddimRepd(Q) and the above description of Z0. □

From this description, we can derive a recursive criterion for the existence of semi-stable representations:

Theorem 5.10.

Let us assume ExtA(T,T)=0. A dimension type (s,d)N×NQ0 is semi-stable if and only if γTs,d=d and there exists no HN type (s,d)*=((s1,d1),,(sr,dr)) with weight (s, d) and sr0 such that  (sn,dn),(sl,dl)A[T]=0   ( for all 1n<lr).

Proof.

Let H:={(s,d)*:rN>1,(s,d)*=((s1,d1),,(sr,dr)) is of HN‐type with weight (s,d)and sr0}.

Obviously H is a finite set. Using Theorem 5.6 we get Rep(s,d)=Rep(s,d)sst (s,d)*HRep(s,d)*HN.

If Rep(s,d)sst, Rep(s,d)sst is of equal dimension like Rep(s,d). So codimRep(s,d)Rep(s,d)*HN>0 for (s,d)*H implies Rep(s,d)sst.

Let n < l. Since (sn,dn) resp. (sl,dl) are semi-stable, we find semi-stable representations N˜ resp. M˜ of A[T] with dim¯ N˜=(sn,dn) resp. dim¯ M˜=(sl,dl) and by using Corollaries 3.6 and 5.7 we get the relation (sn,dn),(sl,dl)A[T]=dimHomA[T](N˜,M˜)dimExtA[T](N˜,M˜).

From μ(sn,dn)>μ(sl,dl) we deduce HomA[T](N˜,M˜)=0, in particular (sn,dn),(sl,dl)A[T]=dimExtA[T](N˜,M˜).

So the equation n<l(sn,dn),(sl,dl)A[T]=0

is fulfilled if and only if (sn,dn),(sl,dl)A[T]=0 for all 1n<lr. □

Example 5.1.

We carry on with Example 2.1 here. Obviously the G(3,1)-orbit of T=(k3[1 0 0]k) is dense in Rep(3,1)(Q). Therefore we have ExtkQ(T,T)=0.

  1. Applying the recursive criterion we derive that (2, 4, 1) is semi-stable. In fact, by using the first criterion in Theorem 5.2, we see that in this case the semi-stability notion equals to the stability notion. From the geometry of the moduli (Theorem 5.8) we can conclude dimM(2,4,1)st(A[T])=4.

  2. Analogous statements as in 1) hold for the dimension vector (3, 6, 2). And we can deduce dim M(3,6,2)st(A[T])=6.

6 Generating semi-invariants

In this section we assume that Q is an acyclic quiver. We determine a set of functions generating the ring of semi-invariants for given dimension vector (s,d)NQ̂0 on Rep(s,d)(A[T]) under the G(s,d) base change action.

We start with a general observation:

Lemma 6.1.

Let G be a linear reductive group and X an affine G-variety. Let AX be a closed and G-stable subset. Then, for every semi-invariant function f:Ak there is a semi-invariant f˜:Xk such that f˜|A=f holds.

Proof.

Let χ be a character of G. We consider the action of G on X×k given by g.(x,λ):=(gx,χ(g)λ), where (x,λ)X×k,gG.

Since A×kX×k is closed and G-stable, the categorical quotient (A×k)//G(X×k)//G is closed, i.e. k[X×k]Gk[A×k]G,ff|A×k is surjective. □

In the following, to simplify the notation we denote by  the path algebra of the one-point extended quiver Q̂.

Let N˜ be a representation of Q̂ with projective resolution 0vQ̂0Âevb(v)θvQ̂0Âeva(v)N˜0.

For M˜Rep(s,d)(Q̂) we apply the functor HomÂ(?,M˜) to this resolution and get 0HomÂ(N˜,M˜)HomÂ(vQ̂0Âeva(v),M˜)HomÂ(θ,M˜)HomÂ(vQ̂0Âevb(v),M˜)ExtÂ(N˜,M˜)0.

The condition dim¯ N˜,(s,d)Q̂=0 is equivalent to vQ̂0a(v)d̂v=vQ̂0b(v)d̂v, that is, in this case we end up with a linear map between vector spaces of equal dimension. Schofield and Van den Bergh proved [Citation13] that all semi-invariant functions arise as linear combination of functions ωN˜:Rep(s,d)(Q̂)k,M˜det(HomÂ(θ,M˜)) induced by representations N˜ of Q̂ with dim¯ N˜,(s,d)Q̂=0.

Using the relation ÂA[T] we can conclude with Lemma 6.1:

Lemma 6.2.

The functions ωN˜ for representations N˜ of A[T] such that dim¯ N˜,(s,d)Q̂=0 generate the ring of invariants on Rep(s,d)(A[T]).

We can improve this description further by using the canonical exact sequence for one-point extensions and the explicit description of the standard projective resolution in Theorem 3.3: Let N˜ be a representation of A[T]. Then we have the canonical exact sequence

which in more detail reads

Obviously N˜0 can be interpreted as a representation of Q, with standard resolution of length 1. Thus, we arrive at the following commutative diagram with exact columns and rows:

Thus, if dim¯ N˜,(s,d)Q̂=0 and ωN˜0 hold, we can conclude from the diagram that the determinants ωN˜0 and ωN˜full can be formed. This discussion shows:

Theorem 6.3.

The ring of semi-invariant functions on Rep(s,d)(A[T]) is generated by the functions ω=ωL·ωN˜ induced by representations L of Q such that dim¯ L,dQ=0 and full representations N˜ of A[T] such that dim¯ N˜,(s,d)Q̂=0.

From the homological properties we can further conclude:

Theorem 6.4.

For a character χ of G(s,d), a representation M˜=(M,V,f:TkVM)Rep(s,d)(A[T]) is χ-semi-stable iff there is a non-trivial finite-dimensional representation N˜=(N,W,g:TkWN) of A[T] such that

  1. HomA[T](N˜,M˜)=0, and

  2. dim¯N˜,(s,d)A[T]=dimW·α:ijαQ1dimTi·dj.

Proof.

Take the standard resolution as described as in Theorem 3.3 P(N˜)N˜0 and consider the cochain complex C:=HomA[T](P(N˜),M˜): C:0HomR˜(N˜,M˜)ϕ̂HomA(ΩAAN,M)HomR(TkW,M)HomA(ΩAA(TkW),M)0.

Then you have Hi(C)=ExtA[T]i(N˜,M˜)(i=0,1,2).

In this way, we achieve the relation: dimHomA[T](N˜,M˜)dimExtA[T](N˜,M˜)+dimExtA[T]2(N˜,M˜)=dimHomR˜(N˜,M˜)dimHomA(ΩAAN,M)dimHomR(TkW,M)+dimHomA(ΩAA(TkW),M).

Both conditions in the theorem are equivalent to ϕ̂ being an isomorphism. □

Example 6.1.

We carry on with Example 5.1 here. Long calculations yields to following generating and algebraic independent semi-invariant functions:

  1. The regular maps Rep(2,4,1)full(A[T])k defined by

    0:(A,B,C,M)det(ABC0(MA0)(0MA)),

    1:(A,B,C,M)det(A|B), 2:(A,B,C,M)det(A|C),

    3:(A,B,C,M)det(A+C|B), 4:(A,B,C,M)det(A|B+C), and 5:(A,B,C,M)det(A+C|B+C) give rise to the geometric quotient π̂:Rep(2,4,1)st(A[T])P4 by Gd in the following way

    π̂:=(01,02,03,04,05).

    Thus, we have M(2,4,1)st(A[T])=P4.

  2. In the case of the dimension vector (3, 6, 2) the regular maps

    Rep(3,6,2)full(A[T])k defined by

    0:(A,B,C,M)det(MAMAMAMAMAMAACBBAC),

    1:(A,B,C,M)det(A|B), 2:(A,B,C,M)det(A|C),

    3:(A,B,C,M)det(A+C|B), 4:(A,B,C,M)det(A+B|C),

    5:(A,B,C,M)det(A|B+C), 6:(A,B,C,M)det(A+B|B+C), and 7:(A,B,C,M)det(A+C|B+C) give rise to the geometric quotient π̂:Rep(3,6,2)st(A[T])P6 by Gd in the following way

    π̂:=(01,02,03,04,05,06,07),

    This shows that M(3,6,2)st(A[T])=P6.

7 Higher Gel’fand MacPherson correspondence

We first recall the definition of quiver Grassmannians (see for example [Citation1]). For a quiver Q, a representation X of Q of dimension vector d and another dimension vector ed, we define GrQe(X) as the set of subrepresentations U of X of dimension vector de. This carries a natural scheme structure as the geometric quotient by the base change group Gd of the set HomQ(X,e)e of surjections (that is, rank e maps) from X to a representation of dimension vector e.

From now on, we assume EndA(T)=k, ExtA(T,T)=0 and γTs,d=d.

In this case, the regular map ϕ:Rep(s,d)full(A[T])HomQ(Ts,d)d,(M,f:TsM)f, which is always a locally trivial fiber bundle [Citation3, Lemma 1.2.], has single element fibers, and thus is an isomorphism of varieties. Moreover, we have the Gd-bundle κ:HomQ(Ts,d)dGrQd(Ts),(θi)iQ0ker(iQ0θi).

All in all, we achieve a Gd-bundle ψ:Rep(s,d)sst(A[T])GrQd,sst(Ts), where GrQd,sst(Ts):=κ°ϕ(Rep(s,d)sst(A[T]))GrQd(Ts) is open.

We thus have an induced map π:GrQd,sst(Ts)M(s,d)sst(A[T]).

The linear reductive group AutQ(Ts) acts naturally on GrQd(Ts) and for [M˜]M(s,d)sst(A[T]) clearly we have π1([M˜])=AutQ(Ts).[(M,f)].

We thus find:

Theorem 7.1

(Higher Gel’fand MacPherson correspondence). There is an isomorphism of varieties: M(s,d)sst(A[T])GrQd,sst(Ts)/AutQ(Ts).

Proof.

By Theorem 5.8 M(s,d)sst(A[T]) is smooth, in particularly normal. Theorem 4.8 shows that the quiver Grassmanian of a representation without self-extensions is irreducible. Since π is surjective the claim follows from [Citation7, Theorem 4.2]. □

8 Motive of the moduli space

In this section, we assume ExtA(T,T)=0 and γTs,d=d.

As an application of the arguments in Theorem 5.9 and the explicit recursive formula given there, we derive a formula for the motive of the (smooth) moduli space M(s,d)sst(C):=M(s,d)sst(A[T]) (over C).

To achieve this, we will follow closely the strategy of [Citation8, Section 6], but replace counts of rational points over finite fields by motives as in the proof of [Citation10, Theorem 3.5]. We denote by K0(Var/C) the free abelian group generated by representatives [X] of isomorphism classes of complex varieties X, modulo the relation [X]=[C]+[U] if C is isomorphic to a closed subvariety of X with open complement isomorphic to U. Multiplication in K0(Var/C) is given by [X]·[Y]=[X×Y]. We denote by L the class of the affine line; the following calculations will be performed in the localization K=K0(Var/C)[L1,(1Ln)1:n1].

At this point, we recall a notation from the Section 5.4. Let H:={(s,d)*:rN>1,(s,d)*=((s1,d1),,(sr,dr)) is of HN‐typewith weight(s,d)and sr0}.

In this section, we write Rep(s,d) for Rep(s,d)full(A[T]) and M(s,d)sst for M(s,d)sst(A[T]).

8.1 Motive

Obviously H is finite and by Theorem 5.6 we have Rep(s,d)=Rep(s,d)sst (s,d)*HRep(s,d)*HN and thus [Rep(s,d)sst]=[Rep(s,d)](s,d)*H[Rep(s,d)*HN] in K. We then find [Rep(s,d)sst][G(s,d)]=[Rep(s,d)][G(s,d)](s,d)*H[Rep(s,d)*HN][G(s,d)].

Fix (s,d)*H. As in the proof of Theorem 5.9 we have Rep(s,d)*HNG(s,d)×P(s,d)*Z0,[Z0] =Ln<lα:ijdildjn+slt,dnQ·i=1r[Rep(di,si)sst], and we arrive at [Rep(s,d)*HN]=[G(s,d)][P(s,d)*]·Ln<lα:ijdildjn+slt,dnQ·i=1r[Rep(di,si)sst].

This provides us with the motivic HN-recursion:

Theorem 8.1.

[Rep(s,d)sst][G(s,d)]=[Rep(s,d)][G(s,d)](s,d)*HLn<lα:ijdildjn+slt,dnQ[Rep(s,d)*]·i=1r[Rep(di,si)sst]  .

Using the arguments of [Citation8, Theorem 6.7.], we see that the following relation iZdimCHi(M(s,d)sst,Q)Li/2=[M(s,d)sst] holds and we obtain:

Theorem 8.2.

Let (s, d) be a dimension vector such that semi-stability and stability coincide. Then, with PG(s,d):=G(s,d)/k× we have: iZdimCHi(M(s,d)sst,Q)Li/2=[Rep(s,d)][PG(s,d)](L1)(s,d)*H1[P(s,d)*]·Ln<lα:ijdildjn+slt,dnQ·i=1r[Rep(di,si)sst]  .

Proof.

As in [Citation8, Proposition 6.6.] we have [M(s,d)sst]=[Rep(s,d)sst][PG(s,d)], and the claim follows from the previous theorem. □

8.2 Applications and examples

Lemma 8.3.

Let Q be of Dynkin type, let Is(Q,d) be the set theoretic quotient of Repd(Q) by the structure group Gd via the base change action, and for MRepd(Q) let HomAepi(T,M):={fHomA(T,M):fissurjective}.

Then, we have [Rep(s,d)]=TsM[M]Is(Q,d),[OM]·[HomAepi(Ts,M)].

Proof.

We consider the map Rep(s,d){MRepd(Q):TsM},(M,f)M. and note that [HomAepi(T,?)] is constant along orbits. □

Overall, we find:

Theorem 8.4.

Let Q be of Dynkin type, and let (s, d) be a dimension vector such that semi-stability and stability coincide. Then we have: iZdimCHi(M(s,d)sst,Q)Li/2=1[PG(s,d)]·TsM[M]Is(Q,d),[OM]·[HomAepi(Ts,M)](L1)·(s,d)*H1[P(s,d)*]·Ln<lα:ijdildjn+slt,dnQ·i=1r[Rep(di,si)sst]=:S(s,d)*  .

Example 8.1.

Let Q=(12),T=(k3[1,0,0]k), and (s,d)=(2,4,1). Then H consists of (1,2,0)>μ(1,2,1),(1,1,1)>μ(1,3,0),(1,1,0)>μ(1,3,1).

We calculate all components in the formula for each element in H: S(1,2,0)>μ(1,2,1)=(L31)(L3L)(L1)4,S(1,1,1)>μ(1,3,0)=1(L1)2,S(1,1,0)>μ(1,3,1)=(L31)L(L1)3.

We continue to calculate and get [Rep(s,d)][PG(s,d)]=L2(L41)(L1)2+(L31)(L41)(L1)(L21)(L2L).

In total, we end up with: iZdimCHi(M(s,d)sst,Q)Li/2=L2(L41)(L1)2+(L31)(L41)(L1)(L21)(L2L)(L31)(L3L)(L1)31(L1)(L31)L(L1)2=1+L+L2+L3+L4.

This result was to be expected if one remembers our explicit calculations of the moduli space.

Acknowledgments

The authors are supported by the DFG SFB/Transregio 191 “Symplektische Strukturen in Geometrie, Algebra und Dynamik.”

References

  • Cerulli Irelli, G., Feigin, E., Reineke, M. (2012). Quiver grassmannians and degenerate flag varieties. Algebra Number Theory 6(1):165–194.
  • Cohn, P. M. (2002). Further Algebra and Applications. London: Springer.
  • Crawley-Boevey, W. (1996). On homomorphisms from a fixed representation to a general representation of a quiver. Trans. Amer. Math. soc. 348(5):1909–1919. DOI: 10.1090/S0002-9947-96-01586-3.
  • Kraft, H., Wiedemann, A. (1985). Geometrische methoden in der invariantentheorie. Wiesbaden: Springer.
  • Le Bruyn, L. (2007). Noncommutative Geometry and Cayley-Smooth Orders. Boca Raton, FL: Chapman and Hall/CRC.
  • Mustaţă, M. (2009). An irreducibility criterion. http://www-personal.umich.edu/∼mmustata/Note1_09.pdf, Last accessed on 2022-05-26.
  • Popov, V. L., Vinberg, E. B. (1994). Invariant theory. In: Algebraic Geometry IV. Berlin: Springer, pp. 123–278.
  • Reineke, M. (2003). The Harder-Narasimhan system in quantum groups and cohomology of quiver moduli. Invent. Math. 152(2):349–368.
  • Reineke, M. (2008). Moduli of representations of quivers. arXiv preprint arXiv:0802.2147.
  • Reineke, M., Stoppa, J., Weist, T. (2012). Mps degeneration formula for quiver moduli and refined gw/kronecker correspondence. Geom. Topol. 16(4):2097–2134. DOI: 10.2140/gt.2012.16.2097.
  • Ringel, C. M. (1984). Integral quadratic forms. In: Tame Algebras and Integral Quadratic Forms. Berlin: Springer, pp. 1–40.
  • Schofield, A. (1992). General representations of quivers. Proc. London Math. Soc. 3(1):46–64.
  • Schofield, A., Van den Bergh, M. (2001). Semi-invariants of quivers for arbitrary dimension vectors. Indag. Math. 12(1):125–138.