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

Motivic (Representation) Stability of Representation Varieties and Character Stacks

Márton Hablicsek Jesse Vogel
Abstract

In this paper, we introduce the notions of motivic representation stability that is an algebraic counterpart of the notion of representation stability. In the process, we also introduce the notion of motivic decomposition for varieties equipped with an action of a finite group GG. This motivic decomposition decomposes the virtual class of the variety with respect to irreducible rational representations of GG.

We also formulate conjectures on motivic representation stability in the context of representation varieties and character stacks, and we verify the conjectures for groups whose virtual classes have been extensively studied.

Keywords: representation stability, Grothendieck ring of varieties, representation variety, character stack.

AMS Subject Classification: 14F45, 14M35, 20C05.

1 Introduction

The purpose of this paper is twofold. On one hand, we provide a framework and computational tools to explore the concept of motivic stability and motivic representation stability, focusing on representation varieties of surface groups, free groups, and free abelian groups. Motivic representation stability extends the notion of representation stability [4] by analyzing virtual classes of varieties in Grothendieck rings, offering an algebraic counterpart to representation stability that simultaneously captures representation stability in various cohomology theories.

On the other hand, we expand Chapter 7 of [26] and formulate conjectures regarding motivic stability and motivic representation stability in the context of representation varieties and character stacks over an algebraically closed field kk of characteristic 0. Specifically, we analyze the following cases.

  • \blacksquare

    Surface Groups, i.e., examining the motivic stability of GG-representation varieties, RepG(Mg)\operatorname{Rep}_{G}(M_{g}), of compact surfaces of increasing genus. In this case, we conjecture the following.

    Conjecture A.

    Let GG be a connected linear algebraic group over kk. Let RepG(Mg)\operatorname{Rep}_{G}(M_{g}) denote the GG-representation variety of the surface group of a smooth compact genus gg surface. Then,

    limg[RepG(Mg)][G2g]=[G/[G,G]][G]\lim_{g\to\infty}\frac{[\operatorname{Rep}_{G}(M_{g})]}{[G^{2g}]}=\frac{[G/[G,G]]}{[G]} (1)

    in the completed Grothendieck rings of stacks, K0(Stckk)^\widehat{\textup{K}_{0}(\textup{\bf{}Stck}_{k})}.

  • \blacksquare

    Free Groups, i.e., examining the motivic representation stability of the representation varieties Gn:=HomGp(Fn,G)G^{n}:=\textup{Hom}_{Gp}(F^{n},G) corresponding to the free groups FnF_{n}. In this case, we conjecture that the character stack is motivically representation stable.

    Conjecture B.

    Let GG be a connected linear algebraic group over kk. Then, the sequence of representation varieties, GnG^{n}, and the sequence of their corresponding character stacks, [Gn/G][G^{n}/G], are motivically representation stable.

  • \blacksquare

    Free Abelian Groups, i.e., examining the motivic representation stability of the representation varieties Cn(G):=HomGp(n,G)C_{n}(G):=\textup{Hom}_{Gp}(\mathbb{Z}^{n},G). We conjecture that sequence Cn(G)C_{n}(G) also satisfies motivic representation stability.

    Conjecture C.

    Let GG be a connected linear algebraic group over kk, and assume that GG has connected center. Then, the sequence of representation varieties Cn(G)C_{n}(G) and their corresponding character stacks are motivically representation stable.

  • \blacksquare

    Stability with variation of the rank for free Abelian groups: in this case, we conjecture that sequence Cn(GLr(k))C_{n}(\textup{GL}_{r}(k)) satisfies motivic stability as rr tends to infinity.

    Conjecture D.

    Fix a positive integer nn. Then, the sequence of representation varieties Cn(GLr(k))C_{n}(\textup{GL}_{r}(k)) and their corresponding character stacks are motivically stable.

We note that Conjectures C and D can be thought of the algebraic versions of Theorem 1.1 and Theorem 9.6 of [20].

Conjectures A, B and C are verified in the paper in the cases of linear algebraic groups for which virtual classes of representation varieties have been studied and thus enough techniques have been developed [9, 13, 25]. These groups include the groups of GLr\textup{GL}_{r}, SLr\textup{SL}_{r} and the groups of upper triangular matrices.

The paper is organized as follows. In Section 2, we provide a framework for motivic representation stability. To achieve this goal, we study motivic decompositions with respect to finite group actions that is a decomposition of a GG-virtual class with respect to the irreducible rational representation of a finite group GG. In Section 3, we study Conjecture A via point-counting methods and Topological Quantum Field theories. These methods are connected via natural transformations [10]. In Section 4, we investigate Conjectures B and C in the cases of GLr\textup{GL}_{r} and SLr\textup{SL}_{r}. We also provide computational tools to find the virtual classes of representation varieties and character stacks corresponding to free or free Abelian groups. In Section 5, we prove Conjecture D in the case of n=2n=2.

2 Preliminaries

In this section, we revisit the notion of motivic representation stability of [26]. The key tool used in this section is a motivic decomposition theorem with respect to rational representations.

2.1 Motivic stability in the Grothendieck rings of GG-varieties

Let SS be a variety over an algebraically closed field kk of characteristic 0. In this paper, we will work with GG-varieties, namely, varieties XX over SS equipped with an action of a linear algebraic group GG over kk so that 1) the map XSX\to S is GG-equivariant (with the trivial action of GG on SS) and 2) XX can be covered by GG-equivariant open affines. Morphisms of GG-varieties are morphisms of varieties over SS that are GG-equivariant.

Definition 2.1.

The Grothendieck ring of GG-varieties, K0(VarSG)\textup{K}_{0}(\textup{\bf{}Var}^{G}_{S}), is the free Abelian group generated by isomorphism classes of GG-varieties over SS with modulo the relations of the form [X]=[U]+[Z][X]=[U]+[Z] where ZZ is a GG-invariant closed subvariety of XX and UU is the corresponding GG-invariant open complement.

Remark 2.2.

Note that if XX and YY are GG-varieties over SS, then X×SYX\times_{S}Y is equipped with a natural GG-action, namely, the diagonal GG-action, making K0(VarSG)\textup{K}_{0}(\textup{\bf{}Var}^{G}_{S}) a ring.

In the case where GG is the trivial group, we obtain the Grothendieck ring of varieties over SS, that we will denote by K0(VarS)\textup{K}_{0}(\textup{\bf{}Var}_{S}). For a GG-variety XX over kk, the class [X]K0(VarkG)[X]\in\textup{K}_{0}(\textup{\bf{}Var}^{G}_{k}) is called the motivic (or virtual) class of XX. We denote the virtual class of 𝔸1\mathbb{A}^{1} (with the trivial GG-action) by qq.

Let GG be a finite group and HH a subgroup of GG. On the level of representations, we have induction and restriction functors. Corresponding to these functors, we have maps on the corresponding Grothendieck rings.

Definition 2.3.

Let GG be a finite algebraic group over kk and HH a subgroup of GG. Let SS be a variety over kk. We define the restriction functor

ResHG:VarSGVarSH\operatorname{Res}_{H}^{G}:\textup{\bf{}Var}_{S}^{G}\to\textup{\bf{}Var}_{S}^{H}

as the functor that regards a GG-variety an HH-variety under the inclusion, and we define the induction functor

IndHG:VarSHVarSG\operatorname{Ind}_{H}^{G}:\textup{\bf{}Var}_{S}^{H}\to\textup{\bf{}Var}_{S}^{G}

as the functor that maps an HH-variety XX over SS to the GG-variety (G×X)/H(G\times X)/H (here HH acts diagonally on G×XG\times X). The resulting variety is indeed a GG-variety with action given by multiplication on the factor of GG.

It is easy to see that these functors descend to the Grothendieck ring of varieties providing maps of Abelian groups

ResHG:K0(VarSG)K0(VarSH)andIndHG:K0(VarSH)K0(VarSG).\operatorname{Res}_{H}^{G}:\textup{K}_{0}(\textup{\bf{}Var}_{S}^{G})\to\textup{K}_{0}(\textup{\bf{}Var}_{S}^{H})\quad\mbox{and}\quad\operatorname{Ind}_{H}^{G}:\textup{K}_{0}(\textup{\bf{}Var}_{S}^{H})\to\textup{K}_{0}(\textup{\bf{}Var}_{S}^{G}).

A little bit more is true. The restriction map ResHG:K0(VarSG)K0(VarSH)\operatorname{Res}_{H}^{G}:\textup{K}_{0}(\textup{\bf{}Var}_{S}^{G})\to\textup{K}_{0}(\textup{\bf{}Var}_{S}^{H}) is a ring homomorphism and it can be defined for any morphism of linear algebraic groups HGH\to G. However, for general linear algebraic groups, more care is needed in the case of the induction functor. In fact, we may face three kinds of problems: 1) the GIT quotient appearing in the induction functor may not exist, 2) the GIT quotient may not be a variety, and 3) the GIT quotient may not be motivic. However, in the case when HH is a closed subgroup of a linear algebraic group GG, all these issues are resolved. Indeed, in this case, the quotient (G×X)/H(G\times X)/H exists [19], and since the action of HH is free on G×XG\times X, the GIT quotient is motivic (see, for instance, Theorem 4.2.11 of [9]). As a result of the discussion above, we have the following.

Proposition 2.4.

Let GG and GG^{\prime} be linear algebraic groups, and ρ:GG\rho:G^{\prime}\to G be a homomorphism of algebraic groups over kk. Let SS be a variety over kk. Then, the restricting the action provides a functor

ResGG:VarSGVarSG\operatorname{Res}_{G^{\prime}}^{G}:\textup{\bf{}Var}_{S}^{G}\to\textup{\bf{}Var}_{S}^{G^{\prime}}

that descends to a map of rings

ResGG:K0(VarSG)K0(VarSG).\operatorname{Res}_{G^{\prime}}^{G}:\textup{K}_{0}(\textup{\bf{}Var}_{S}^{G})\to\textup{K}_{0}(\textup{\bf{}Var}_{S}^{G^{\prime}}).

Furthermore, let HH be a closed subgroup of GG. Then, the induction functor

IndHG:VarSHVarSG\operatorname{Ind}_{H}^{G}:\textup{\bf{}Var}_{S}^{H}\to\textup{\bf{}Var}_{S}^{G}

defined by sending an HH-variety XX over SS to the variety (G×X)/H(G\times X)/H descends to a map of Abelian groups

IndHG:K0(VarSH)K0(VarSG).\operatorname{Ind}_{H}^{G}:\textup{K}_{0}(\textup{\bf{}Var}_{S}^{H})\to\textup{K}_{0}(\textup{\bf{}Var}_{S}^{G}).

2.1.1 Motivic stability

In this paper, we are concerned with families of varieties and their limiting virtual class. Explicitly, we consider the localization MqG:=K0(VarkG)[q1]M_{q}^{G}:=\textup{K}_{0}(\textup{\bf{}Var}^{G}_{k})[q^{-1}]. This ring has a natural increasing filtration given by the powers of qq:

0FnMqGFn+1MqGMqG0\subseteq\dots\subseteq F_{n}M_{q}^{G}\subseteq F_{n+1}M_{q}^{G}\subseteq\dots\subseteq M_{q}^{G}

where FnMqGF_{n}M_{q}^{G} is the subgroup generated by the elements of MqGM_{q}^{G} of the form [X]qs\frac{[X]}{q^{s}} where XX is an irreducible variety of dimension at most s+ns+n. We denote the completion of MqGM_{q}^{G} with respect to this filtration by MqG^\widehat{M_{q}^{G}}.

The ring MqG^\widehat{M_{q}^{G}} is equipped with a topology coming from the completion that allows us to consider limits of families of virtual classes as in [23].

Definition 2.5.

We say that a family of GG-varieties {Xn}n\{X_{n}\}_{n} is motivically stable if the limit

limn[Xn]qdimXn\lim_{n\to\infty}\frac{[X_{n}]}{q^{\dim X_{n}}}

exists in MqG^\widehat{M_{q}^{G}}.

It is easy to see that the restriction and induction maps respect the filtrations, meaning that if HH is a closed subgroup of a linear algebraic group GG over kk, then

ResHG(FnMqG)FnMqH\operatorname{Res}_{H}^{G}(F_{n}M_{q}^{G})\subseteq F_{n}M_{q}^{H}

and

IndHG(FnMqH)Fn+cMqG\operatorname{Ind}_{H}^{G}(F_{n}M_{q}^{H})\subseteq F_{n+c}M_{q}^{G}

where cc is the codimension of HH in GG. Therefore, we obtain the following.

Corollary 2.6.

Let HH be a closed subgroup of a linear algebraic group GG over kk. Then, the restriction and induction maps provide a continuous group homomorphism:

ResHG:MqG^MqH^andIndHG:MqH^MqG^.\operatorname{Res}_{H}^{G}:\widehat{M_{q}^{G}}\to\widehat{M_{q}^{H}}\quad\mbox{and}\quad\operatorname{Ind}_{H}^{G}:\widehat{M_{q}^{H}}\to\widehat{M_{q}^{G}}.

In particular, we have that

  • \blacksquare

    if the family of GG-varieties {Xn}n\{X_{n}\}_{n} is motivically stable, then the family of HH-varieties {ResHG(Xn)}n\{\operatorname{Res}_{H}^{G}(X_{n})\}_{n} is also motivically stable,

  • \blacksquare

    if the family of HH-varieties {Yn}n\{Y_{n}\}_{n} is motivically stable, then the family of GG-varieties {IndHG(Yn)}n\{\operatorname{Ind}_{H}^{G}(Y_{n})\}_{n} is also motivically stable.

In the context of motivic stability, one of the most important sequence of varieties that has been studied is the sequence of symmetric powers {SymGn(X)}n\{\operatorname{Sym}^{n}_{G}(X)\}_{n} of GG-varieties: the nn-th symmetric power of a variety XX, SymGnX:=Xn/Sn\operatorname{Sym}^{n}_{G}X:=X^{n}/S_{n}, is naturally equipped with a GG-variety structure induced by the diagonal action.

The following lemma is key in order to establish motivic stability.

Lemma 2.7 (Proposition 4.2 in [23]).

Let XX be a GG-variety and ZXZ\subset X a closed GG-invariant subvariety of small dimension: dimZ<dimX\dim Z<\dim X. Then, the symmetric powers of XX stabilize if and only if the symmetric powers of the open complement U=XZU=X\setminus Z stabilize. Moreover, we have

limnSymGn(X)qndimX=ZG(Z,qdimX)limnSymGn(U)qndimU\lim_{n\to\infty}\frac{\operatorname{Sym}_{G}^{n}(X)}{q^{n\dim X}}=Z_{G}(Z,q^{-\dim X})\lim_{n\to\infty}\frac{\operatorname{Sym}_{G}^{n}(U)}{q^{n\dim U}}

where ZG(Z,qdimX)Z_{G}(Z,q^{-\dim X}) denotes the motivic zeta function of the GG-variety XX in the sense of [15].

2.1.2 Motivic stability in the Grothendieck ring of stacks

In Section 3 a slightly different motivic stability will be considered. For that, we consider Ekedahl’s version of the Grothendieck ring of stacks defined as follows [5].

Definition 2.8.

The Grothendieck ring of stacks K0(Stckk)\textup{K}_{0}(\textup{\bf{}Stck}_{k}) is defined as the Abelian group generated by stacks of finite type over kk with affine stabilizers module the relations 1) [𝔛]=[]+[𝔘][\mathfrak{X}]=[\mathfrak{Z}]+[\mathfrak{U}] where \mathfrak{Z} is a closed substack of 𝔛\mathfrak{X} with open complement 𝔘\mathfrak{U} and 2) the relations of the form

[𝔈]=[𝔸kn×𝔛][\mathfrak{E}]=[\mathbb{A}^{n}_{k}\times\mathfrak{X}]

for every vector bundle 𝔈𝔛\mathfrak{E}\to\mathfrak{X} of rank nn.

Ekedahl shows that the Grothendieck ring of stacks over kk is isomorphic to the localization of the Grothendieck ring of varieties, K0(Vark)\textup{K}_{0}(\textup{\bf{}Var}_{k}), by inverting the class of the affine line qq and the classes of the form qn1q^{n}-1.

We define motivic stability in this ring parallel to the case of the Grothendieck ring of varieties. Namely, we consider the natural increasing filtration given by the powers of the symbols qt1q^{t}-1:

0FnK0(Stckk)Fn+1K0(Stckk)K0(Stckk)0\subseteq\dots\subseteq F_{n}K_{0}(\textup{\bf{}Stck}_{k})\subseteq F_{n+1}K_{0}(\textup{\bf{}Stck}_{k})\subseteq\dots\subseteq K_{0}(\textup{\bf{}Stck}_{k})

where FnK0(Stckk)F_{n}K_{0}(\textup{\bf{}Stck}_{k}) is the subgroup generated by the elements of K0(Stckk)K_{0}(\textup{\bf{}Stck}_{k}) of the form [X](qi1)si\frac{[X]}{\prod(q^{i}-1)^{s_{i}}} where XX is an irreducible variety, the product in the denominator is finite and XX is of dimension at most s+isis+\sum i\cdot s_{i}. We denote the completion of K0(Stckk)K_{0}(\textup{\bf{}Stck}_{k}) with respect to this filtration by K0(Stckk)^\widehat{\textup{K}_{0}(\textup{\bf{}Stck}_{k})}.

Remark 2.9.

The E-polynomial is a motivic measure of K0(Var)\textup{K}_{0}(\textup{\bf{}Var}_{\mathbb{C}}) that sends a smooth and projective variety XX to E(X):=i,jhi,juivj[u,v]E(X):=\sum_{i,j}h^{i,j}u^{i}v^{j}\in\mathbb{Z}[u,v], and sends the affine line to E(𝔸1):=uvE(\mathbb{A}^{1}):=uv. Using the above, we show that the E-polynomial defines a motivic measure of K0(Stck)^\widehat{\textup{K}_{0}(\textup{\bf{}Stck}_{\mathbb{C}})}. Indeed, consider the ring [u,v]\mathbb{Z}[u,v], and invert the element uvu\cdot v in this ring. Then, we have a natural filtration given by the powers of 1uv\frac{1}{uv} and one can consider the completion of the localized ring with respect to this filtration, [u,v,1uv]^\widehat{\mathbb{Z}[u,v,\frac{1}{uv}]}. Now, for the virtual class of a stack of the form [V/GLn][V/\textup{GL}_{n}], we define its E-polynomial as

E(V)(uv)n2i=1n(1+1(uv)i+1(uv)2i+)\frac{E(V)}{(uv)^{n^{2}}}\prod_{i=1}^{n}(1+\frac{1}{(uv)^{i}}+\frac{1}{(uv)^{2}i}+...)

that is well-defined in [u,v,1uv]^\widehat{\mathbb{Z}[u,v,\frac{1}{uv}]}. Note, that the E-polynomial does not depend on the representation [V/GLn][V/\textup{GL}_{n}] [16, 1]. We can see that the E-polynomial respects the filtration on K0(Stck)\textup{K}_{0}(\textup{\bf{}Stck}_{\mathbb{C}}), thus it provides a motivic measure K0(Stck)^[u,v,1uv]^\widehat{\textup{K}_{0}(\textup{\bf{}Stck}_{\mathbb{C}})}\to\widehat{\mathbb{Z}[u,v,\frac{1}{uv}]}.

2.2 Motivic representation stability

The goal of this section is to provide a framework in studying representation stability via virtual classes. Since, the proofs of the representation stability results [20] for representation varieties, character varieties and character stacks do not depend on the actual cohomology theory, it is expected that representation stability holds on the motivic level.

The difficulty in approaching this problem using the Grothendieck ring of varieties is twofold. First, it is not clear how to decompose a variety with an SnS_{n}-action into subvarieties corresponding to representations of SnS_{n}. Second, in order to do any kind of computations, one needs that this decomposition satisfies certain properties, for instance, Künneth-formuala, etc.

These difficulties cannot be solved. To avoid the first problem, we have to make a choice on how to decompose a variety according to the representations of SnS_{n}. The choice is basically an appropriate set of subgroups in SnS_{n}. To deal with the second problem, we will show that, even though not every SnS_{n}-varieties can be used in Künneth-type formulas, there are still sufficient SnS_{n}-varieties that can be used for computational purposes.

We begin by addressing the first problem. The representations of SnS_{n} are parametrized by Young tableaux: for a partition λ\lambda of nn, we denote the corresponding representation by VλV_{\lambda}. For a general partition λ=(λ1,λ2,)\lambda=(\lambda_{1},\lambda_{2},...) (with λ1λ2\lambda_{1}\geq\lambda_{2}\geq...) and for an integer n|λ|+λ1:=2λ1+λ2+n\geq|\lambda|+\lambda_{1}:=2\lambda_{1}+\lambda_{2}+..., we denote the partition of nn given by (n|λ|,λ1,λ2,)(n-|\lambda|,\lambda_{1},\lambda_{2},...) by λ[n]\lambda[n]. In representation stability, one is interested in the stability of the representations of Vλ[n]V_{\lambda[n]}.

Corresponding to a partition λ=(λ1,λ2,)\lambda=(\lambda_{1},\lambda_{2},...) of nn, we have a subgroup

Sλ:=Sλ1×Sλ2×Sn.S_{\lambda}:=S_{\lambda_{1}}\times S_{\lambda_{2}}\times...\leq S_{n}.

Thus, for the sequence of representations Vλ[n]V_{\lambda[n]}, we obtain a sequence of subgroups Sλ[n]SnS_{\lambda[n]}\leq S_{n}. After setting up the notation, we arrive at the definition of motivic representation stability.

Definition 2.10.

Let {Xn}n\{X_{n}\}_{n} be a sequence of varieties over kk with an action of G×{Sn}nG\times\{S_{n}\}_{n} (with GG being a fixed linear algebraic group). We say that this sequence is motivically representation stable if the sequences of varieties [Xn/Sλ[n]][X_{n}/S_{\lambda[n]}] are motivically stable in MqG^\widehat{M_{q}^{G}} for all partitions λ\lambda.

In many of our cases, GG will be the trivial group. The following observation is key to understanding motivic representation stability.

Example 2.11.

Let XX be a GG-variety, and consider the sequence Xn=XnX_{n}=X^{n} with the natural G×SnG\times S_{n}-action (where SnS_{n} permutes the coordinates). Then, we see that

[Xn/Sλ[n]]=SymGn|λ|X×iSymGλiX[X^{n}/S_{\lambda[n]}]=\operatorname{Sym}_{G}^{n-|\lambda|}X\times\prod_{i}\operatorname{Sym}_{G}^{\lambda_{i}}X

meaning that the sequence XnX^{n} is motivically representation stable if and only if the sequence SymGnX\operatorname{Sym}^{n}_{G}X is motivically stable in MqG^\widehat{M_{q}^{G}}.

In the rest of the section, we motivate why we believe that Definition 2.10 is a right definition for motivic representation stability in the context of the Grothendieck ring of varieties.

2.3 Motivic decomposition with respect to representations

Let GG be a finite group and consider, R(G)R_{\mathbb{Q}}(G), the representation ring of GG over \mathbb{Q} that is the Grothendieck group of finite dimensional \mathbb{Q}-representations of GG where the ring structure is induced by the tensor product of representations. In this section, we wish to decompose a GG-variety over a variety SS with respect to the rational representations of GG, in other words, we wish to construct a ring map

K0(VarSG)K0(VarS)ZR(G)\textup{K}_{0}(\textup{\bf{}Var}_{S}^{G})\to\textup{K}_{0}(\textup{\bf{}Var}_{S})\otimes_{Z}R_{\mathbb{Q}}(G)

so that for a variety XX over SS the part that corresponds to the trivial representation is [X/G][X/G] (indeed, the tensor product on the right-hand side has a natural structure of a ring, because both K0(VarS)\textup{K}_{0}(\textup{\bf{}Var}_{S}) and R(G)R_{\mathbb{Q}}(G) are rings). We denote the desired motivic decomposition by [X]GK0(VarS)R(G)[X]_{G}\in\textup{K}_{0}(\textup{\bf{}Var}_{S})\otimes_{\mathbb{Z}}R_{\mathbb{Q}}(G) for a GG-variety XX.

This goal will not be achieved in general. The main difficulties arise from a) the lack of a motivic Künneth-formula, see Example 3.6.17. of [26], and b) over-constraints coming from subgroups of GG, see Example 3.6.9 of [26].

The representation ring, R(G)R_{\mathbb{Q}}(G), is equipped with the standard inner product ,\langle,\rangle that is given on two irreducible representations, V1V_{1}, V2V_{2}, by V1,V2=0\langle V_{1},V_{2}\rangle=0 if the irreducible representations are not isomorphic and V1,V2=1\langle V_{1},V_{2}\rangle=1, if the representations are isomorphic. This inner product can be extended to an inner product

,:K0(VarS)ZR(G)K0(VarS)ZR(G)K0(VarS)\langle,\rangle:\textup{K}_{0}(\textup{\bf{}Var}_{S})\otimes_{Z}R_{\mathbb{Q}}(G)\otimes_{\mathbb{Z}}\textup{K}_{0}(\textup{\bf{}Var}_{S})\otimes_{Z}R_{\mathbb{Q}}(G)\to\textup{K}_{0}(\textup{\bf{}Var}_{S})

by [X1]V1,[X2]V2=[X1][X2]V1,V2.\langle[X_{1}]\otimes V_{1},[X_{2}]\otimes V_{2}\rangle=[X_{1}][X_{2}]\otimes\langle V_{1},V_{2}\rangle.

Definition 2.12.

Let GG be a finite group, and let \mathcal{H} be a set of conjugacy classes of subgroups of GG. We define the map

Ψ:R(G)[H]\Psi_{\mathcal{H}}:R_{\mathbb{Q}}(G)\to\bigoplus_{[H]\in\mathcal{H}}\mathbb{Z}

as [V]TH,ResHG(V)[H][V]\mapsto\langle T_{H},\operatorname{Res}_{H}^{G}(V)\rangle_{[H]\in\mathcal{H}}. We say that a set of conjugacy classes of subgroups \mathcal{H} is good if the map

Ψ:R(G)[H]\Psi_{\mathcal{H}}\otimes\mathbb{Q}:R_{\mathbb{Q}}(G)\otimes_{\mathbb{Z}}\mathbb{Q}\to\bigoplus_{[H]\in\mathcal{H}}\mathbb{Q}

is an isomorphism of vector spaces.

Remark 2.13.

The reason we only consider conjugacy classes of subgroups is that both TH,ResHG(V)\langle T_{H},\operatorname{Res}_{H}^{G}(V)\rangle and [X/H][X/H] only depend on the conjugacy class of the subgroup HH.

Remark 2.14.

We note that the inner product on R(G)R_{\mathbb{Q}}(G) naturally extends to an inner product

,:R(G)R(G)\langle,\rangle:R_{\mathbb{Q}}(G)\otimes_{\mathbb{Z}}\mathbb{Q}\otimes_{\mathbb{Q}}R_{\mathbb{Q}}(G)\otimes_{\mathbb{Z}}\mathbb{Q}\to\mathbb{Q}

on the vector space R(G)R_{\mathbb{Q}}(G)\otimes_{\mathbb{Z}}\mathbb{Q}.

Our key observation is that good sets of conjugacy classes of subgroups provide motivic decompositions.

Lemma 2.15.

Let \mathcal{H} be a good set of conjugacy classes of subgroups. Then, there exists a unique map of vector spaces

K0(VarSG)K0(VarS)R(G)\textup{K}_{0}(\textup{\bf{}Var}_{S}^{G})\otimes_{\mathbb{Z}}\mathbb{Q}\to\textup{K}_{0}(\textup{\bf{}Var}_{S})\otimes_{\mathbb{Z}}R_{\mathbb{Q}}(G)\otimes_{\mathbb{Z}}\mathbb{Q}

so that for any GG-variety XX, the image of [X]G:=[X]1[X]_{G}:=[X]\otimes 1 satisfies

TH,ResHG[X]G=[X/H]\langle T_{H},\operatorname{Res}_{H}^{G}[X]_{G}\rangle=[X/H]

in K0(VarS)\textup{K}_{0}(\textup{\bf{}Var}_{S})\otimes_{\mathbb{Z}}\mathbb{Q} for all [H][H]\in\mathcal{H}.

Proof.

Given a GG-variety XX, consider the potential element [X]G=V[UV][V]λVK0(VarS)R(G)[X]_{G}=\sum_{V}[U_{V}]\otimes[V]\otimes\lambda_{V}\in\textup{K}_{0}(\textup{\bf{}Var}_{S})\otimes_{\mathbb{Z}}R_{\mathbb{Q}}(G)\otimes_{\mathbb{Z}}\mathbb{Q} (where the summation goes over the irreducible rational representations VV of GG). In order to satisfy our assumption, we need that

V[UV]λVTH,V=[X/H]1\sum_{V}[U_{V}]\otimes\lambda_{V}\langle T_{H},V\rangle=[X/H]\otimes 1

holds for every [H][H]\in\mathcal{H}. Since \mathcal{H} is a good set of conjugacy classes of subgroups, this equation system has a unique solution proving our statement.

In many cases, natural choices exist for good sets of conjugacy classes of subgroups.

Example 2.16.

Let GG be a finite cyclic group /n\mathbb{Z}/n\mathbb{Z}. In this case, consider the set of all subgroups \mathcal{H}. It is easy to see that in this case, the number of subgroups and the number of rational representations agree (namely the number of divisors of nn), and, thus, the corresponding map Ψ\Psi_{\mathcal{H}}\otimes\mathbb{Q} is an isomorphism.

Example 2.17.

Let G1G_{1} and G2G_{2} be finite groups and 1\mathcal{H}_{1} and 2\mathcal{H}_{2} be good sets of conjugacy classes of subgroups of G1G_{1} and G2G_{2} respectively. Then, 1×2\mathcal{H}_{1}\times\mathcal{H}_{2} provides a good set of conjugacy classes of subgroups for G1×G2G_{1}\times G_{2}. This implies that any finite Abelian group has a choice of a good set of conjugacy classes of subgroups, i.e for any finite Abelian group, a motivic decomposition exists. This set and, thus, the decomposition depend on the choice of the decomposition of the Abelian group into cyclic factors.

Example 2.18.

Consider the symmetric group SnS_{n}, and for each partition λ\lambda the subgroup Sλ:=Sλ1×Sλ2×S_{\lambda}:=S_{\lambda_{1}}\times S_{\lambda_{2}}\times.... Consider \mathcal{H} to be the set of conjugacy classes of the SλS_{\lambda}. Denote by VλV_{\lambda} the irreducible representation corresponding to λ\lambda. Now, for [V]=aλ[Vλ]R(Sn)[V]=\sum a_{\lambda}[V_{\lambda}]\in R_{\mathbb{Q}}(S_{n}), we have

ΨSn(V)=(TSλ,ResSλSnV)Sλ=(IndSλSn(TSλ),V)Sλ=(μaμKμ,λ)Sλ\Psi_{S_{n}}^{\mathcal{H}}(V)=\left(\langle T_{S_{\lambda}},\operatorname{Res}_{S_{\lambda}}^{S_{n}}V\rangle\right)_{S_{\lambda}\in\mathcal{H}}=\left(\langle\operatorname{Ind}_{S_{\lambda}}^{S_{n}}(T_{\mathrm{S}_{\lambda}}),V\rangle\right)_{S_{\lambda}\in\mathcal{H}}=\left(\sum_{\mu}a_{\mu}K_{\mu,\lambda}\right)_{S_{\lambda}\in\mathcal{H}}

where Kμ,λK_{\mu,\lambda} are the Kostka numbers (see [7] Corollary 4.39). Since, Kμ,μ=1K_{\mu,\mu}=1 and Kμ,λ=0K_{\mu,\lambda}=0 for μ<λ\mu<\lambda (with respect to the lexigraphic ordering), we have that ΨSn(V)\Psi_{S_{n}}^{\mathcal{H}}(V) is an isomorphism.

The example above motivates Definition 2.10, in fact, the following holds.

Theorem 2.19 (Proposition 7.3.3 of [26]).

Let {Xn}n\{X_{n}\}_{n} be a sequence of varieties over kk equipped with action of the symmetric groups {Sn}n\{S_{n}\}_{n}. Then, {Xn}n\{X_{n}\}_{n} is motivically representation stable, if and only if, for any partition λ\lambda, in the motivic decomposition, the coefficient of [Xn]Sn[X_{n}]_{S_{n}} of Vλ[n]1V_{\lambda[n]}\otimes 1 is motivically stable. In other words, the motivic decomposition is motivically stable.∎

2.3.1 Künneth-formula

The vector space K0(VarSG)R(G)\textup{K}_{0}(\textup{\bf{}Var}_{S}^{G})\otimes_{\mathbb{Z}}R_{\mathbb{Q}}(G)\otimes_{\mathbb{Z}}\mathbb{Q} has a natural ring structure coming from the ring structures of K0(VarS)\textup{K}_{0}(\textup{\bf{}Var}_{S}) and R(G)R_{\mathbb{Q}}(G) and \mathbb{Q}. In this section, we investigate Künneth-tpye formulae for motivic decompositions.

We show that although the K”unneth-formula does not hold in general for the motivic decomposition, it holds for simple cases. To show this, we start with an easy lemma.

Lemma 2.20 (Lemma 3.6.18. of [26]).

Let GG be a finite algebraic group, \mathcal{H} a good set of conjugacy classes of subgroups of GG. Let 𝒱GK0(VarSG)\mathcal{V}_{G}^{\mathcal{H}}\subseteq\textup{K}_{0}(\textup{\bf{}Var}^{G}_{S}) be the subset of elements [X][X] so that [XY]G=[X]G[Y]G[XY]_{G}=[X]_{G}[Y]_{G}. Then, 𝒱G\mathcal{V}_{G}^{\mathcal{H}} is a K0(VarS)\textup{K}_{0}(\textup{\bf{}Var}_{S})-subalgebra of K0(VarSG)\textup{K}_{0}(\textup{\bf{}Var}_{S}^{G}).∎

This lemma enables us to provide a large set of varieties for which the motivic Künneth-formula holds.

Theorem 2.21.

[Theorem 3.6.19. of [26]] Let GG be a finite algebraic group over a field kk so that kk contains all |G||G|-th roots of unity, \mathcal{H} a good set of conjugacy classes of subgroups of GG. Then,

  • \blacksquare

    If GG acts linearly on 𝔸1\mathbb{A}^{1}, then 𝒱G\mathcal{V}_{G}^{\mathcal{H}} contains the class of 𝔸1\mathbb{A}^{1}.

  • \blacksquare

    If GG acts diagonally on 𝔸n\mathbb{A}^{n}, then 𝒱G\mathcal{V}_{G}^{\mathcal{H}} contains the class of 𝔸n\mathbb{A}^{n}.

  • \blacksquare

    If GG acts diagonally on n\mathbb{P}^{n}, then 𝒱G\mathcal{V}_{G}^{\mathcal{H}} contains the class of n\mathbb{P}^{n}.

  • \blacksquare

    The discreet group scheme GG with the natural translation action by GG lies in 𝒱G\mathcal{V}_{G}^{\mathcal{H}}.

Proof.

We provide an alternative short proof for the first statement, then the second and third statement follows from the first one from the lemma before.

Since, [𝔸1/H]=[𝔸1][\mathbb{A}^{1}/H]=[\mathbb{A}^{1}] for any linear action of a group on [𝔸1][\mathbb{A}^{1}], we have that the motivic decomposition of [𝔸1][\mathbb{A}^{1}] is just [𝔸1]TG[\mathbb{A}^{1}]\otimes T_{G}. Thus, to show the first statement, we need to show that [𝔸1×Y/H]=[𝔸1][Y/H][\mathbb{A}^{1}\times Y/H]=[\mathbb{A}^{1}][Y/H] for all YK0(VarSG)Y\in\textup{K}_{0}(\textup{\bf{}Var}_{S}^{G}) and all [H]fH[H]\in fH.

Consider, NHN\leq H, the kernel of the representation HGL1(k)H\to\textup{GL}^{1}(k) corresponding to the linear action of HH on 𝔸1\mathbb{A}^{1}. Then,

[𝔸1×Y/H]=[(𝔸1×Y/N)/(H/N)]=[𝔸1][(Y/N)/(H/N)]=[𝔸1][Y/H[\mathbb{A}^{1}\times Y/H]=[(\mathbb{A}^{1}\times Y/N)/(H/N)]=[\mathbb{A}^{1}][(Y/N)/(H/N)]=[\mathbb{A}^{1}][Y/H

showing that if the statement is true for the smaller group H/NH/N, then it is also true for HH. Therefore, using induction, we can assume that HH is a cyclic group acing freely on 𝔾m=𝔸1{0}\mathbb{G}_{m}=\mathbb{A}^{1}\setminus\{0\}. In this case, [𝔾m×Y/H][\mathbb{G}_{m}\times Y/H] is a 𝔾m\mathbb{G}_{m}-torsor over [Y/H][Y/H], so

[𝔾m×Y/H]=[𝔾m][Y/H][\mathbb{G}_{m}\times Y/H]=[\mathbb{G}_{m}][Y/H]

impyling the proof as [𝔸1×Y/H]=[𝔾m×Y/H]+[Y/H][\mathbb{A}^{1}\times Y/H]=[\mathbb{G}_{m}\times Y/H]+[Y/H]. ∎

3 Motivic Stability of Representation Varieties of Surface groups

Let MgM_{g} be a compact genus gg smooth surface, and GG a linear algebraic group. In this section, we are interested whether the varieties RepG(Mg):=HomGp(π1(Mg),G)\operatorname{Rep}_{G}(M_{g}):=\textup{Hom}_{Gp}(\pi_{1}(M_{g}),G) are motivically stable in terms of gg. A priori, HomGp(π1(Mg),G)\textup{Hom}_{Gp}(\pi_{1}(M_{g}),G) is only endowed with a set structure, however,

HomGp(π1(Mg),G)={A1,B1,,Ag,BgG2g|i=1g[Ai,Bi]=id}\textup{Hom}_{Gp}(\pi_{1}(M_{g}),G)=\{A_{1},B_{1},...,A_{g},B_{g}\in G^{2g}|\prod_{i=1}^{g}[A_{i},B_{i}]=\textup{id}\}

and thus RepG(Mg)\operatorname{Rep}_{G}(M_{g}) can, indeed, be realized as a closed subvariety of G2gG^{2g}.

Explicitly, we are interested whether the limit

limg[RepG(Mg)][G2g]\lim_{g\to\infty}\frac{[\operatorname{Rep}_{G}(M_{g})]}{[G^{2g}]}

exists in the completed Grothendieck ring of stacks, K0(Stckk)^\widehat{\textup{K}_{0}(\textup{\bf{}Stck}_{k})}. Informally, this limit measures the probability that the 2g2g elements A1,B1,,Ag,BgA_{1},B_{1},...,A_{g},B_{g} of the group GG satisfy the property

i=1g[Ai,Bi]=id.\prod_{i=1}^{g}[A_{i},B_{i}]=\textup{id}.

3.1 Finite field heuristics

We expect that as the genus gets larger and larger the product of commutators i=1g[Ai,Bi]\prod_{i=1}^{g}[A_{i},B_{i}] spreads evenly across the commutator subgroup [G,G][G,G]. In fact, this happens in the case of finite groups GG. In this case, using a classical result of Frobenius [6], RepG(Mg)\operatorname{Rep}_{G}(M_{g}) is a finite set of cardinality

#RepG(Mg)=#Gχ(#Gχ(1))2g2.\#\operatorname{Rep}_{G}(M_{g})=\#G\sum_{\chi}\left(\frac{\#G}{\chi(1)}\right)^{2g-2}. (2)

Thus, in the case of finite groups, the limit 1 exists, namely

limg#RepG(Mg)#G2g=limg1#Gχ1χ(1)2g2=1[G,G].\lim_{g\to\infty}\frac{\#\operatorname{Rep}_{G}(M_{g})}{\#G^{2g}}=\lim_{g\to\infty}\frac{1}{\#G}\sum_{\chi}\frac{1}{\chi(1)^{2g-2}}=\frac{1}{[G,G]}.

The above computation suggests that such limit should exist motivically in the case of connected linear algebraic groups over \mathbb{C}.

3.2 The case of reductive groups with connected centers

In this section, we show that the limit 1 exists in the context of the motivic measure given by the E-polynomial (see Remark 2.9) for reductive groups GG with connected center.

The GG-representation varieties of surfaces, RepG(Mg)\operatorname{Rep}_{G}(M_{g}), are PORC count (polynomial on residue classes) [2] meaning that #RepG(𝔽q)(Mg)\#\operatorname{Rep}_{G(\mathbb{F}_{q})}(M_{g}) is a polynomial, qG,g(t)[t]q_{G,g}(t)\in\mathbb{Q}[t], in terms of qq supposing that q1q\equiv 1 module d(G)d(G^{\vee}), the modulus of the Langlands dual group. Using a result of Katz, this implies that the E-polynomial of the RepG(Mg)\operatorname{Rep}_{G}(M_{g}) is given as qG,g(uv)q_{G,g}(uv).

Furthemore, the summand of the right-hand side of Equation 2 corresponding to 1-dimensional representations

#G#(G/[G,G])(#G)2g2\#G\cdot\#(G/[G,G])\cdot(\#G)^{2g-2}

is also PORC count, showing the sum of the rest of the summands together is also PORC count. Using Remark 5.2 of [2], the rest of the summands of the form (Gχ(1))\left(\frac{G}{\chi(1)}\right) are of smaller dimensions than GG, so their contribution is negligible in the limiting E-polynomial (since we raise these pieces to the 2g2g-th power).

To summarize the above, we have the following.

Corollary 3.1.

Let GG be a reductive group over \mathbb{C} with connected center. Then,

limgE(RepG(Mg)E(G)2g=E(G/[G,G])E(G).\lim_{g\to\infty}\frac{E(\operatorname{Rep}_{G}(M_{g})}{E(G)^{2g}}=\frac{E(G/[G,G])}{E(G)}.

3.3 General linear algebraic groups

We conjecture, based on the finite field heuristics and Corollary 3.1, that the limit 1 exists.

Conjecture 3.2.

Let GG be a connected linear algebraic group over kk. Let RepG(Mg)\operatorname{Rep}_{G}(M_{g}) denote the GG-representation variety of the surface group of a smooth compact genus gg surface. Then,

limg[RepG(Mg)][G2g]=[G/[G,G]][G]\lim_{g\to\infty}\frac{[\operatorname{Rep}_{G}(M_{g})]}{[G^{2g}]}=\frac{[G/[G,G]]}{[G]}

in the completed Grothendieck rings of stacks, K0(Stckk)^\widehat{\textup{K}_{0}(\textup{\bf{}Stck}_{k})}.

We verify the conjecture in two main cases: 1. in the case of G=SL2(k)G=\textup{SL}_{2}(k) and in the case of upper triangular matrix groups GG of rank up to 5. In fact, the virtual classes of GG-representation varieties of surface groups are notoriously difficult to compute, namely, only the cases above have been solved successfully [9, 13, 25]. In all such cases, the computations were performed using Topological Quantum Field Theories.

Example 3.3 (The case of upper triangular matrices).

Explicit computations have been done for the group of upper triangular matrices or unipotent matrices of ranks 2, 3, 4 and 5 [13, 25]:

  • \blacksquare

    [Rep𝕌2(Mg)]=q2g1(q1)2g+1((q1)2g1+1)=q2g1(q1)4g+l.o.t[\operatorname{Rep}_{\mathbb{U}_{2}}(M_{g})]=q^{2g-1}(q-1)^{2g+1}\left((q-1)^{2g-1}+1\right)=q^{2g-1}(q-1)^{4g}+l.o.t

  • \blacksquare

    [Rep𝕌3(Mg)=q6g3(q1)6g+l.o.t[\operatorname{Rep}_{\mathbb{U}_{3}}(M_{g})=q^{6g-3}(q-1)^{6g}+l.o.t

  • \blacksquare

    [Rep𝕌4(Mg)=q12g6(q1)8g+l.o.t[\operatorname{Rep}_{\mathbb{U}_{4}}(M_{g})=q^{12g-6}(q-1)^{8g}+l.o.t

  • \blacksquare

    [Rep𝕌5(Mg)]=q20g10(q1)10g+l.o.t[\operatorname{Rep}_{\mathbb{U}_{5}}(M_{g})]=q^{20g-10}(q-1)^{10g}+l.o.t

providing the proof of Conjecture 3.2 for the low rank upper triangular matrix groups.

Example 3.4 (The case of SL2(k)\textup{SL}_{2}(k)).

Using the explicit description of eigenvectors and eigenvalues [24] one gets

[RepSL2(k)(Mg)]=12q2g1(q+1)2g1(q1)(22g+q1)+12q2g1(q1)2g1(q+1)(22g+q3)+[\operatorname{Rep}_{\textup{SL}_{2}(k)}(M_{g})]=\frac{1}{2}q^{2g-1}(q+1)^{2g-1}(q-1)(2^{2g}+q-1)+\frac{1}{2}q^{2g-1}(q-1)^{2g-1}(q+1)(2^{2g}+q-3)+
+(q2g1+q)(q1)2g1(q+1)2g1+(q^{2g-1}+q)(q-1)^{2g-1}(q+1)^{2g-1}

showing

limg[RepSL2(k)(Mg)][SL2(k)]2g=1q(q1)(q+1)=1[SL2(k),SL2(k)]\lim_{g\rightarrow\infty}\frac{[\operatorname{Rep}_{\textup{SL}_{2}(k)}(M_{g})]}{[\textup{SL}_{2}(k)]^{2g}}=\frac{1}{q(q-1)(q+1)}=\frac{1}{[\textup{SL}_{2}(k),\textup{SL}_{2}(k)]}

in K0(Stckk)^\widehat{\textup{K}_{0}(\textup{\bf{}Stck}_{k})}.

3.4 Counterexamples via non-connected groups

The assumption that the linear algebraic group is connected is crucial in Conjecture 3.2. In fact, in the case of non-connected groups, Lang’s theorem [17] fails, and the finite field heuristics fail. To give an example, we consider G=𝔾m/2G=\mathbb{G}_{m}\rtimes\mathbb{Z}/2\mathbb{Z}, where the action of /2\mathbb{Z}/2\mathbb{Z} on 𝔾m\mathbb{G}_{m} is given by xx1x\mapsto x^{-1}. In [11], the authors described the virtual class of the corresponding representation variety as

[RepG(Mg)]=(q1)2g1(q3+22g+1)[\operatorname{Rep}_{G}(M_{g})]=(q-1)^{2g-1}(q-3+2^{2g+1})

providing

limg[RepG(Mg)][G]2g=2q11[G,G]=1q1\lim_{g\rightarrow\infty}\frac{[\operatorname{Rep}_{G}(M_{g})]}{[G]^{2g}}=\frac{2}{q-1}\neq\frac{1}{[G,G]}=\frac{1}{q-1}

in K(Stck/k)\textup{K}(\textup{\bf{}Stck}/k).

The problem is even more severe if one considers the corresponding character stack. In fact, in [11], the class of the character stack is computed as

[𝔛G(Mg)])=(q1)2g2(22g+1+q3)2+(q+1)2g2(22g+1+q1)2.[\mathfrak{X}_{G}(M_{g})])=\frac{\left(q-1\right)^{2g-2}\left(2^{2g+1}+q-3\right)}{2}+\frac{\left(q+1\right)^{2g-2}\left(2^{2g+1}+q-1\right)}{2}.

Using this computation, it is clear that the limit in Conjecture 3.2 does not exist.

4 Motivic representation stability of representation varieties of free groups and free Abelian groups

Goal of the section is Motivic representation stability of representation varieties of free groups and free Abelian groups

4.1 Free groups

In this section, we study the motivic representation stability of the representation varieties Gn:=HomGp(Fn,G)G^{n}:=\textup{Hom}_{Gp}(F^{n},G) corresponding to the free groups FnF_{n} on nn letters, and of the corresponding character stacks [Gn/G][G^{n}/G] where the group GG acts by conjugation on GnG^{n}. The symmetric groups, SnS_{n}, act by permuting the coordinates of GnG^{n}, and the conjugation action by the group GG commutes with the action of SnS_{n}.

First, we show that the representation varieties are motivically representation stable which follows from the following easy lemma.

Lemma 4.1.

Let XX be a variety over kk, so that [X][X] is a polynomial in qq in K0(Vark)\textup{K}_{0}(\textup{\bf{}Var}_{k}). Then, the symmetric powers, Symn(X)\operatorname{Sym}^{n}(X), are motivically stable if and only if [X][X] is a monic polynomial of qq.

Proof.

We compute the motivic zeta function of XX from the polynomial form [X]=i=0saiqi[X]=\sum_{i=0}^{s}a_{i}q^{i} using that the motivic zeta function is motivic:

Z(X,t)=i=0s(n=0[Symn(qi)]tn)ai.Z(X,t)=\prod_{i=0}^{s}\left(\sum_{n=0}^{\infty}[\operatorname{Sym}^{n}(q^{i})]t^{n}\right)^{a_{i}}.

Using that [Symn(qi)]=qni[\operatorname{Sym}^{n}(q^{i})]=q^{ni} (see for instance, Lemma 4.4 of [12]), we have

Z(X,t)=i=0s(n=0qnitn)ai=i=0s1(1qit)ai.Z(X,t)=\prod_{i=0}^{s}\left(\sum_{n=0}^{\infty}q^{ni}t^{n}\right)^{a_{i}}=\prod_{i=0}^{s}\frac{1}{(1-q^{i}t)^{a_{i}}}.

Thus, if as=1a_{s}=1, then

limn[Symn(X)]qns=((1t)Z(X,t/qs))|t=1=1ti=0s(1qist)ai=1i=0s1(1qist)ai\lim_{n\to\infty}\frac{[\operatorname{Sym}^{n}(X)]}{q^{ns}}=\left((1-t)Z(X,t/q^{s})\right)|_{t=1}=\frac{1-t}{\prod_{i=0}^{s}(1-q^{i-s}t)^{a_{i}}}=\frac{1}{\prod_{i=0}^{s-1}(1-q^{i-s}t)^{a_{i}}}

and if as>1a_{s}>1, it is easy to see that the limit does not exist. ∎

This lemma above shows that the GG-representation varieties corresponding to free groups where GG is a connected linear algebraic group are motivically representation stable proving Conjecture B using Example 2.11. Indeed, GnG^{n} is motivically representation stable if and only if SymnG\operatorname{Sym}^{n}G is motivically stable as a sequence of GG-varieties.

Now, we turn our attention to the character stacks of free groups. Again, by Example 2.11, GnG^{n} is motivically representation stable if and only if SymGnG\operatorname{Sym}^{n}_{G}G is motivically stable as a sequence of GG-varieties.

The key examples we consider (similarly to the other examples in the paper) are G=GLrG=\textup{GL}_{r} and the groups of upper triangular matrices. The key statement we need is the following.

Proposition 4.2.

Consider the natural action of GLr\textup{GL}_{r} on 𝔸r\mathbb{A}^{r}. Then,

limn[SymGLrn(𝔸r)]qnr=[GLr]i=1r(qrqi)\lim_{n\to\infty}\frac{[\operatorname{Sym}^{n}_{\textup{GL}_{r}}(\mathbb{A}^{r})]}{q^{nr}}=\frac{[\textup{GL}_{r}]}{\prod_{i=1}^{r}(q^{r}-q^{i})}

in MqG^\widehat{M_{q}^{G}} where GLr\textup{GL}_{r} acts free on [GLr][\textup{GL}_{r}] by left multiplication.

Proof.

Let us denote by XnX_{n} the part of Symn(𝔸r)\operatorname{Sym}^{n}(\mathbb{A}^{r}) where GLr\textup{GL}_{r} acts freely. We claim that

limn[Xn]qnr=limn[Symn(𝔸r)]qnr=1\lim_{n\to\infty}\frac{[X_{n}]}{q^{nr}}=\lim_{n\to\infty}\frac{[\operatorname{Sym}^{n}(\mathbb{A}^{r})]}{q^{nr}}=1

holds in Mq^\widehat{M_{q}}. The claim implies the statement as GLr\textup{GL}_{r} is a special group, because in this case

limn[Xn]qnr=limn[Xn/GLr][GLr]qnr=[GLr]i=1r(qrqi)\lim_{n\to\infty}\frac{[X_{n}]}{q^{nr}}=\lim_{n\to\infty}\frac{[X_{n}/\textup{GL}_{r}][\textup{GL}_{r}]}{q^{nr}}=\frac{[\textup{GL}_{r}]}{\prod_{i=1}^{r}(q^{r}-q^{i})}

holds in MqG^\widehat{M_{q}^{G}} since the GG-action on [Xn/GLr][X_{n}/\textup{GL}_{r}] is trivial, so [Xn/GLr][X_{n}/\textup{GL}_{r}] can be replaced by [Symn(𝔸r)]GLr\frac{[\operatorname{Sym}^{n}(\mathbb{A}^{r})]}{\textup{GL}_{r}}. In order to show the claim we will show that XnX_{n} can be covered by varieties of negligible dimension.

In fact, we have a cover of the form

σSn{A,(x1,,xn)(GLrid)×𝔸rn|i:Axi=xσ(i)}.\bigsqcup_{\sigma\in S_{n}}\{A,(x_{1},...,x_{n})\in(\textup{GL}_{r}\setminus id)\times\mathbb{A}^{rn}|\forall i:Ax_{i}=x_{\sigma(i)}\}.

This cover is of dimension at most dimGLr+nrn\dim\textup{GL}_{r}+nr-n, because either 1) all xix_{i}’s have to be eigenvectors of AA with eigenvalue 1 or 2) at least one of the equations xi=Axjx_{i}=Ax_{j} hold with iji\neq j, so xix_{i} is determined by xjx_{j}. This concludes the proof of the statement. ∎

This proposition implies an important technical statement that was used in [20] in the context of FI-modules.

Corollary 4.3.

Let GG be a linear algebraic group acting on 𝔸r\mathbb{A}^{r} via a homomorphism ϕ:GGLr\phi:G\to\textup{GL}_{r}. Then,

limn[SymGn(𝔸r)]qnr=[GLr]i=1r(qqqi)\lim_{n\to\infty}\frac{[\operatorname{Sym}^{n}_{G}(\mathbb{A}^{r})]}{q^{nr}}=\frac{[\textup{GL}_{r}]}{\prod_{i=1}^{r}(q^{q}-q^{i})}

holds in MqG^\widehat{M_{q}^{G}} where GG acts on GLr\textup{GL}_{r} via the homomorphism ϕ\phi.

Proof.

We have that ResHG\operatorname{Res}_{H}^{G} provides a continuous map MqG^MqH^\widehat{M_{q}^{G}}\to\widehat{M_{q}^{H}} using Corollary 2.6. Now, the statement follows from the realization that

ResGGLrSymGLrn=SymGnResGGLr.\operatorname{Res}^{\textup{GL}_{r}}_{G}\circ\operatorname{Sym}^{n}_{\textup{GL}_{r}}=\operatorname{Sym}^{n}_{G}\circ\operatorname{Res}^{\textup{GL}_{r}}_{G}.

Using the corollary above, we are ready to show that the GG-character stacks corresponding to free groups satisfy motivic representation stability in the cases of G=GLrG=\textup{GL}_{r} or G=𝕌rG=\mathbb{U}_{r}.

Theorem 4.4.

The GG-character stacks corresponding to free groups satisfy motivic representation stability in the cases of G=GLrG=\textup{GL}_{r} or G=𝕌rG=\mathbb{U}_{r}.

Proof.

In order to show motivic representation stability, it is enough to show that the sequence of symmetric powers SymGn(X)\operatorname{Sym}^{n}_{G}(X) is motivically stable (see Example 2.11). In the case of G=GLrG=\textup{GL}_{r}, this follows from Proposition 4.2 and Lemma 2.7. In the case of G=𝕌rG=\mathbb{U}_{r}, the upper triangular matrices act on the upper half of the matrices linearly, and thus Proposition 4.2 implies the statement. ∎

4.2 Free Abelian groups

In this section, we study the motivic representation stability of the representation varieties Cn(G):=HomGp(n,G)C_{n}(G):=\textup{Hom}_{Gp}(\mathbb{Z}^{n},G) corresponding to the free Abelian groups n\mathbb{Z}^{n}, and of the corresponding character stacks [Cn(G)/G][C_{n}(G)/G]. As before, the symmetric groups, SnS_{n}, act by permuting the coordinates of GnG^{n}, and the group GG acts by conjugation.

4.2.1 Branching matrix

First, we provide a computational technique to find the virtual classes of the variety of commuting nn-tuples, Cn(G)C_{n}(G), for reductive groups via branching matrices [22]. We illustrate the method for the reductive group GL2(k)\textup{GL}_{2}(k).

The virtual class of GL2(k)\textup{GL}_{2}(k) is q(q1)(q21)q(q-1)(q^{2}-1). It has three types of elements

  • \blacksquare

    scalars: the closed subvariety of scalars has virtual class q1q-1.

  • \blacksquare

    JJ-type: the open subvariety of the closed subvariety of matrices with the same eigenvalue (but not scalars) has virtual class (q1)(q21)(q-1)(q^{2}-1). These are the matrices that are conjugate to a matrix of the form

    (λ10λ).\left(\begin{array}[]{cc}\lambda&1\\ 0&\lambda\end{array}\right).
  • \blacksquare

    MM-type: the open subvariety of matrices with two distinct eigenvalues has virtual class (q1)(q3q2q)(q-1)(q^{3}-q^{2}-q). These are the matrices that are conjugate to a matrix of the form

    (λ00μ)\left(\begin{array}[]{cc}\lambda&0\\ 0&\mu\end{array}\right)

    with λμ\lambda\neq\mu.

A different way to calculate the class of MM-type elements is as follows. Any such element is conjugate to a matrix

(λ00μ)\left(\begin{array}[]{cc}\lambda&0\\ 0&\mu\end{array}\right)

and thus

M=GL2(k)/D×{(λ00μ)|λμ}//2M=\textup{GL}_{2}(k)/D\times\left\{\left(\begin{array}[]{cc}\lambda&0\\ 0&\mu\end{array}\right)|\lambda\neq\mu\right\}/\mathbb{Z}/2\mathbb{Z}

where the /2\mathbb{Z}/2\mathbb{Z}-action is given by swapping the eigenvalues λμ\lambda\leftrightarrow\mu and on GL2(k)/D\textup{GL}_{2}(k)/D is given by multiplication by (0110)\left(\begin{array}[]{cc}0&1\\ 1&0\end{array}\right) on the right.

With respect to the /2\mathbb{Z}/2\mathbb{Z}-action, we have the following.

Lemma 4.5.

The virtual classes of the ++ and - parts of the varieties are as follows.

  • \blacksquare

    [C]:=[GL2(k)]+=q2[C]:=[\textup{GL}_{2}(k)]^{+}=q^{2}

  • \blacksquare

    [GL2(k)]=q[\textup{GL}_{2}(k)]^{-}=q

  • \blacksquare

    [{(λ00μ)|λμ,λμ0}]+=q22q+1\left[\left\{\left(\begin{array}[]{cc}\lambda&0\\ 0&\mu\end{array}\right)|\lambda\neq\mu,\lambda\mu\neq 0\right\}\right]^{+}=q^{2}-2q+1

  • \blacksquare

    [{(λ00μ)|λμ,λμ0}]=1q\left[\left\{\left(\begin{array}[]{cc}\lambda&0\\ 0&\mu\end{array}\right)|\lambda\neq\mu,\lambda\mu\neq 0\right\}\right]^{-}=1-q

  • \blacksquare

    [{(λ00μ)|λμ0}]+=q2q\left[\left\{\left(\begin{array}[]{cc}\lambda&0\\ 0&\mu\end{array}\right)|\lambda\mu\neq 0\right\}\right]^{+}=q^{2}-q

  • \blacksquare

    [{(λ00μ)|λμ0}]=1q\left[\left\{\left(\begin{array}[]{cc}\lambda&0\\ 0&\mu\end{array}\right)|\lambda\mu\neq 0\right\}\right]^{-}=1-q.

Proof.

The statements follow after stratifications from the fact that [𝔸n]+=𝔸n[\mathbb{A}^{n}]^{+}=\mathbb{A}^{n} for any linear action. ∎

For the sake of simplicity, we denote by XX the variety GL2(k)/D\textup{GL}_{2}(k)/D. We note that using the quotient map, XX is naturally a variety over CC. The lemma above enables us to compute the virtual class of MM in a different way.

Lemma 4.6.

As a variety over CC, [M][M] is a linear combination of the classes [X][X] and [C][C].

Proof.

Using the motivic Künneth-formula (Theorem 2.21), we have that

M=(GL2(k)×{(λ00μ)|λμ,λμ0})//2=M=(\textup{GL}_{2}(k)\times\left\{\left(\begin{array}[]{cc}\lambda&0\\ 0&\mu\end{array}\right)|\lambda\neq\mu,\lambda\mu\neq 0\right\})/\mathbb{Z}/2\mathbb{Z}=
GL2(k)+{(λ00μ)|λμ,λμ0}++GL2(k){(λ00μ)|λμ,λμ0}.\textup{GL}_{2}(k)^{+}\left\{\left(\begin{array}[]{cc}\lambda&0\\ 0&\mu\end{array}\right)|\lambda\neq\mu,\lambda\mu\neq 0\right\}^{+}+\textup{GL}_{2}(k)^{-}\left\{\left(\begin{array}[]{cc}\lambda&0\\ 0&\mu\end{array}\right)|\lambda\neq\mu,\lambda\mu\neq 0\right\}^{-}.

Thus, the class of MM over CC is

[M]=[C](q22q+1)+([X][C])(1q)=[C](q2q)+[X](1q).[M]=[C](q^{2}-2q+1)+([X]-[C])(1-q)=[C](q^{2}-q)+[X](1-q).

Thus, [M]=q2(q2q)+(q2+q)(1q)=(q2q)(q2q1)[M]=q^{2}(q^{2}-q)+(q^{2}+q)(1-q)=(q^{2}-q)(q^{2}-q-1).

Next, we parametrize the isomorphism classes of centralizer subgroups. In the case of GL2(k)\textup{GL}_{2}(k), we have the following possibilities.

  • \blacksquare

    The whole group GL2(k)\textup{GL}_{2}(k): this is the centralizer of the scalar matrices.

  • \blacksquare

    JJ-type: the centralizer of an element of JJ-type. These centralizers are parametrized by GL2(k)/J\textup{GL}_{2}(k)/J the left cosets of the subgroup J={(μb0μ)}J=\left\{\left(\begin{array}[]{cc}\mu&b\\ 0&\mu\end{array}\right)\right\} via conjugating the centralizer {(λb0λ)|λ0}\left\{\left(\begin{array}[]{cc}\lambda&b\\ 0&\lambda\end{array}\right)|\lambda\neq 0\right\} with elements of GL2(k)/J\textup{GL}_{2}(k)/J.

  • \blacksquare

    MM-type

Next, we construct the branching matrix that encodes how the centralizers of nn-tuples of elements change. In the case of GL2(k)\textup{GL}_{2}(k), this is fairly simple, the centralizer can only change from the whole group to another centralizer.

Explicitly, we have the following.

  • \blacksquare

    The centralizer of an nn-tuple of commuting elements was the whole group GL2(k)\textup{GL}_{2}(k): In this case, the centralizer remains the whole group if and only if the next element is a scalar matrix. The centralizer becomes JJ-type if the next element is of JJ-type, and it becomes MM-type if the next element is of MM-type.

  • \blacksquare

    The centralizer of an nn-tuple of commuting elements was of JJ-type: In this case, the centralizer has to remain the same if we add one more commuting element. Indeed, after conjugation the centralizer is of the form {(λb0λ)}\left\{\left(\begin{array}[]{cc}\lambda&b\\ 0&\lambda\end{array}\right)\right\}, and thus any element of the nn-tuple is of form (μc0μ)\left(\begin{array}[]{cc}\mu&c\\ 0&\mu\end{array}\right) so the next elements must be as well. This provides q(q1)q(q-1) choices.

  • \blacksquare

    The centralizer of an nn-tuple of commuting elements was of MM-type: In this case, the centralizer has to remain the same if we add one more commuting element, moreover the centralizer is conjugate to the subgroup of diagonal matrices, thus the centralizers are parametrized by C=(GL2(k)/D)+C=(\textup{GL}_{2}(k)/D)^{+}. The n+1n+1-st element thus needs to be conjugate to a diagonal matrix by the same group elements. This means that if Cn(M)C_{n}(M) denotes the variety of nn-tuples of commuting elements of MM-type, then Cn+1(M)=Cn(M)×C(C×{(λ00μ)|λμ0})//2C_{n+1}(M)=C_{n}(M)\times_{C}\left(C\times\left\{\left(\begin{array}[]{cc}\lambda&0\\ 0&\mu\end{array}\right)|\lambda\mu\neq 0\right\}\right)/\mathbb{Z}/2\mathbb{Z}.

Now, we focus on computing Cn+1(M)C_{n+1}(M) using the recursion

Cn+1(M)=Cn(M)×C(C×{(λ00μ)|λμ0})//2.C_{n+1}(M)=C_{n}(M)\times_{C}\left(C\times\left\{\left(\begin{array}[]{cc}\lambda&0\\ 0&\mu\end{array}\right)|\lambda\mu\neq 0\right\}\right)/\mathbb{Z}/2\mathbb{Z}.

We have the following easy lemma to help us.

Lemma 4.7.

We have that [X×CX]=2[X]=2q2+2q[X\times_{C}X]=2[X]=2q^{2}+2q.

Proof.

It is easy to see that the /2\mathbb{Z}/2\mathbb{Z}-action is free on XX providing the statement above. ∎

Lemma 4.6 shows that the class [C1(M)]=[M][C_{1}(M)]=[M] over CC is a linear combination of [C][C] and [X][X]. Similarly, the class of M¯=(C×{(λ00μ)|λμ0})//2\overline{M}=\left(C\times\left\{\left(\begin{array}[]{cc}\lambda&0\\ 0&\mu\end{array}\right)|\lambda\mu\neq 0\right\}\right)/\mathbb{Z}/2\mathbb{Z} is also a linear combination of CC and [X][X], namely, we have

[¯M]=[C][q21]+[X](1q).\overline{[}M]=[C][q^{2}-1]+[X](1-q).

The discussion above implies that the branching matrix of GL2(k)\textup{GL}_{2}(k) is of the following form.

Theorem 4.8.

The branching matrix of GL2(k)\textup{GL}_{2}(k) in the basis given by [D][D], [J][J], [C][C] and [X][X] is as follows

A=(q1000q1q(q1)00(q2q)(q1)0q210(1q)(q1)01q(q1)2).A=\left(\begin{array}[]{cccc}q-1&0&0&0\\ q-1&q(q-1)&0&0\\ (q^{2}-q)(q-1)&0&q^{2}-1&0\\ (1-q)(q-1)&0&1-q&(q-1)^{2}\end{array}\right).

Since [M]=(q2q)[C]+(1q)[X][M]=(q^{2}-q)[C]+(1-q)[X] and the virtual classes are [D]=q1[D]=q-1, [J]=(q1)(q21)[J]=(q-1)(q^{2}-1), [C]=q2[C]=q^{2}, [X]=q2+q[X]=q^{2}+q, we have the following.

Theorem 4.9.

The virtual class of the nn-tuples of commuting elements in GL2(k)\textup{GL}_{2}(k) is given by

wAn1vTwA^{n-1}v^{T}

where w=(q1,(q1)(q21),q2,q2+q)w=(q-1,(q-1)(q^{2}-1),q^{2},q^{2}+q) and v=(1,1,q2q,1q)v=(1,1,q^{2}-q,1-q).

This allows us to compute the classes of the nn-tuples of commuting elements explicitly after diagonalizing the matrix AA.

Theorem 4.10.

We have that the virtual class of the nn-tuples of commuting elements in GL2(k)\textup{GL}_{2}(k) is

12q(q21)(2(q2q)n12(q1)n1+(q1)(q21)n1+(q1)2n1).\frac{1}{2}q(q^{2}-1)\left(2(q^{2}-q)^{n-1}-2(q-1)^{n-1}+(q-1)(q^{2}-1)^{n-1}+(q-1)^{2n-1}\right).

In particular, the variety of nn-tuples of commuting elements is of dimension 2n+22n+2.

Remark 4.11.

The above computation can be performed, in principle, for any group with finitely many isomorphism classes of centralizers, thus, for instance, for reductive groups [8].

4.2.2 Motivic representation stability of commuting nn-tuples

Although it is clear from Theorem 4.10 that the varieties of commuting nn-tuples do not satisfy motivic stability, it motivates our conjecture that these varieties satisfy motivic representation stability. In fact, the failure of the motivic stability is a consequence of the non-trivial contribution of the non-trivial representations in the motivic representation stability.

The case of GL2\textup{GL}_{2}: Before explaining motivic representation stability for a general linear reductive algebraic group, we illustrate the general proof in the case of GL2\textup{GL}_{2}. In the previous section, we stratified the variety of commuting triples into three pieces according to the corresponding centralizer groups:

Sn=Cn(D)={A1,,An|Ais are scalars}S^{n}=C_{n}(D)=\{A_{1},...,A_{n}|A_{i}^{\prime}s\mbox{ are scalars}\}
Cn(J)={A1,An|some Ai is conjugate to (λ10λ)}C_{n}(J)=\left\{A_{1},...A_{n}|\mbox{some }A_{i}\mbox{ is conjugate to }\left(\begin{array}[]{cc}\lambda&1\\ 0&\lambda\end{array}\right)\right\}
Cn(M)={A1,An|some Ai is conjugate to (λ00μ)μλ}C_{n}(M)=\left\{A_{1},...A_{n}|\mbox{some }A_{i}\mbox{ is conjugate to }\left(\begin{array}[]{cc}\lambda&0\\ 0&\mu\end{array}\right)\mu\neq\lambda\right\}

By simultaneously conjugating the elements AiA_{i}, we can express Cn(J)C_{n}(J) as

Cn(J)=IndHGL2(JnSn)C_{n}(J)=\operatorname{Ind}_{H}^{\textup{GL}_{2}}(J^{n}\setminus S^{n})

where J={(μb0μ)}J=\left\{\left(\begin{array}[]{cc}\mu&b\\ 0&\mu\end{array}\right)\right\} is the subgroup of upper triangular matrices with equal eigenvalues and H=JH=J is the stabilizer of JJ. Similarly,

Cn(M)=IndKGL2(DnSn)C_{n}(M)=\operatorname{Ind}_{K}^{\textup{GL}_{2}}(D^{n}\setminus S^{n})

where DD is the subgroup of diagonal matrices, KK being the stabilizer of DD. Since the dimension of SnS^{n} is negligible compared to DnD^{n} or JnJ^{n}, it is enough to show that

IndHGL2(Jn)andIndKGL2(Dn)\operatorname{Ind}_{H}^{\textup{GL}_{2}}(J^{n})\quad\mbox{and}\quad\operatorname{Ind}_{K}^{\textup{GL}_{2}}(D^{n})

are motivically representation stable. Since, the action of SnS_{n} and GL2\textup{GL}_{2} commute on the commuting nn-tuples, we have that

IndHGL2(Jn)/Sλ[n]=IndHGL2(SymHn|λ|J×iSymHλiJ)\operatorname{Ind}_{H}^{\textup{GL}_{2}}(J^{n})/\mathrm{S}_{\lambda[n]}=\operatorname{Ind}_{H}^{\textup{GL}_{2}}(\operatorname{Sym}_{H}^{n-|\lambda|}J\times\prod_{i}\operatorname{Sym}_{H}^{\lambda_{i}}J)

and similarly

IndKGL2(Dn)/Sλ[n]=IndKGL2(SymKn|λ|D×iSymKλiD).\operatorname{Ind}_{K}^{\textup{GL}_{2}}(D^{n})/\mathrm{S}_{\lambda[n]}=\operatorname{Ind}_{K}^{\textup{GL}_{2}}(\operatorname{Sym}_{K}^{n-|\lambda|}D\times\prod_{i}\operatorname{Sym}_{K}^{\lambda_{i}}D).

The action of HH (and KK resp.) are linear on JJ viewed as an open subvariety of 𝔸2\mathbb{A}^{2} (and on DD viewed as an open subvariety of 𝔸2\mathbb{A}^{2}), thus motivic representation stability holds using Corollary 4.3.

The case of a general linear reductive algebraic group: The strategy above can be used to show the following general statement.

Theorem 4.12.

Let GG be a linear reductive algebraic group over an algebraically closed field kk. Assume that all the Abelian subgroups of GG that are maximal under inclusion are connected. Then, the variety of commuting nn-tuples in GG satisfies motivic representation stability.

Proof.

In a linear reductive algebraic group over an algebraically closed field kk has finitely many centralizers of the form ZG(g)Z_{G}(g) up to conjugation [8]. The Abelian subgroups that are maximal under inclusion are finite intersections of centralizers of elements, therefore, we have only finitely many maximal Abelian subgroups under conjugation.

Now, using the same idea as in the case of GL2\textup{GL}_{2}, we have that

limn[Cn(G)/Sλ[n]]qdimCn(G)=limnAG:max Abelian[IndNG(A)GL2(An)/Sλ[n]]qdimCn(G).\lim_{n\to\infty}\frac{[C_{n}(G)/S_{\lambda[n]}]}{q^{\dim C_{n}(G)}}=\lim_{n\to\infty}\sum_{A\leq G:\mbox{max Abelian}}\frac{[\operatorname{Ind}_{N_{G}(A)}^{\textup{GL}_{2}}(A^{n})/S_{\lambda[n]}]}{q^{\dim C_{n}(G)}}.

This limit exists in Mq^\widehat{M_{q}} using Lemma 4.1 and Corollary 2.6, because the summation is finite. ∎

We can strengthen the theorem above to deal with the case of character stacks in the case of the general linear group GLr\textup{GL}_{r}.

Corollary 4.13.

The GLr\textup{GL}_{r}-character stacks of commuting nn-tuples of GLr\textup{GL}_{r} satisfy motivic representation stability, i.e, the sequence Cn(GLr)C_{n}(\textup{GL}_{r}). is motivially representation stable in K0(VarkGLr)\textup{K}_{0}(\textup{\bf{}Var}_{k}^{\textup{GL}_{r}}).

Proof.

The proof for GL3\textup{GL}_{3} works similarly to the case of GL2\textup{GL}_{2} by working through all possible centralizers. In the general case, one can use the theorem of [21, 14] stating that the maximal Abelian subgroups of GLr\textup{GL}_{r} are open subsets of affine spaces and the action of GLr\textup{GL}_{r} can be extended to a linear action on the affine spaces. Therefore, Corollary 4.3 implies that the GLr\textup{GL}_{r}-varieties of commuting nn-tuples of GLr\textup{GL}_{r} satisfy motivic representation stability. ∎

Counterexample

The assumption of the connectedness in Theorem 4.12 is crucial.

Proposition 4.14.

The variety of commuting nn-tuples in SLr\textup{SL}_{r} does not satisfy motivic representation stability.

Proof.

The center of SLr\textup{SL}_{r} is not connected, it is isomorphic to the finite cyclic group of order rr. Therefore, the maximal Abelian subgroups are also not connected, their virtual classes are not monic polynomials in qq. Since, the center commutes with any normalizer, the virtual class of the underlying variety of IndNG(A)SLr(An/Sλ[n])\operatorname{Ind}_{N_{G}(A)}^{\textup{SL}_{r}}(A^{n}/S_{\lambda[n]}) is also not a monic polynomial in qq, and thus using Theorem 4.12, Lemma 4.1 shows that the variety of commuting nn-tuples in SLr\textup{SL}_{r} can not satisfy motivic representation stability. ∎

Similar argument works for other linear algebraic groups as well with non-connected centers.

5 Motivic stability of representation varieties of free groups and free Abelian groups corresponding to reductive groups of increasing rank

In this section, we provide an algebraic analogue of Theorem 9.6 of [20]: we conjecture that the family of varieties of commuting nn-tuples in GLr\textup{GL}_{r} satisfies motivic stability as the rank, rr, tends to infinity. To provide evidence, we prove the conjecture for n=2n=2.

Conjecture 5.1.

Fix a positive integer nn. Consider the family of varieties Cn(GLr(k))C_{n}(\textup{GL}_{r}(k)). Then,

limr[Cn(GLr(k))]qdimCn(GLr(k))\lim_{r\to\infty}\frac{[C_{n}(\textup{GL}_{r}(k))]}{q^{\dim C_{n}(\textup{GL}_{r}(k))}}

exists in Mq^\widehat{M_{q}}.

In the case of n=1n=1, the conjecture is immediate. This section is devoted to prove a quantitative version of Conjecture 5.1 in the case n=2n=2.

Theorem 5.2.

Consider the family of varieties, C2(GLr(k))C_{2}(\textup{GL}_{r}(k)). Then,

limr[C2(GLr(k))]qr[GLr(k)]=1\lim_{r\to\infty}\frac{[C_{2}(\textup{GL}_{r}(k))]}{q^{r}[\textup{GL}_{r}(k)]}=1

in K0(Stckk)^\widehat{\textup{K}_{0}(\textup{\bf{}Stck}_{k})}.

The above theorem can be thought of as the motivic version of [18]. We begin with analyzing the virtual classes of all the conjugacy classes of elements.

Theorem 5.3.

Consider the family of groups, GLr(k)\textup{GL}_{r}(k). Then, the virtual class of the family of the union of the conjugacy classes of GLr(k)\textup{GL}_{r}(k) is motivically stable.

Proof.

The proof is a straightforward generalization of the proof in [18], we add the details for the sake of completeness. Consider, GLr\textup{GL}_{r}, and possible characteristic polynomial of an [r×r][r\times r] matrix

p(x)=i=1k(xλi)aip(x)=\prod_{i=1}^{k}(x-\lambda_{i})^{a_{i}}

where the matrix has kk Jordan blocks of sizes a1a2aka_{1}\geq a_{2}\geq...\geq a_{k} with eigenvalues λi\lambda_{i} respectively. Consider the alteration of p(x)p(x),

q(x)=i=1k(1λix)ai.q(x)=\prod_{i=1}^{k}(1-\lambda_{i}x)^{a_{i}}.

We parametrize the different classes of conjugacy classes via factoring q(x)q(x) using partitions corresponding to the size of the Jordan blocks: q(x)=l=1rhl(x)lq(x)=\prod_{l=1}^{r}h_{l}(x)^{l} where hl(x)h_{l}(x) is the product of the factors of q(x)q(x) of the form (1λix)(1-\lambda_{i}x) where ai=la_{i}=l. The polynomials, hl(x)h_{l}(x), are polynomials with constant term 1 and of degree nln_{l} satisfying l=1rlnl=r\sum_{l=1}^{r}ln_{l}=r. Thus, the polynomials, hl(x)h_{l}(x) are parametrized by 𝔸nl1×(𝔸1{0})\mathbb{A}^{n_{l}-1}\times(\mathbb{A}^{1}\setminus\{0\}). Furthermore, the polynomials hl(x)h_{l}(x) uniquely determine Jordan block, therefore, the virtual class of the families of conjugacy classes corresponding to the partition (a1,a2,)(a_{1},a_{2},...) is given as l=1r(qniqni1)\prod_{l=1}^{r}(q^{n_{i}}-q^{{n_{i}}-1}). Using the same calculation as in [18], we get that the virtual class of all conjugacy classes, [𝒞r][\mathcal{C}_{r}], satisfies

limr[𝒞r]qr=1\lim_{r\to\infty}\frac{[\mathcal{C}_{r}]}{q^{r}}=1

in MqG^\widehat{M_{q}^{G}}, proving the statement. ∎

Remark 5.4.

For sake of clarity, we illustrate the computation for GL2(k)\textup{GL}_{2}(k). The Jordan forms come in three different forms.

  • \blacksquare

    The scalar matrices: The virtual class of the family of conjugacy classes of matrices that are conjugate to scalar matrices is q1q-1.

  • \blacksquare

    The JJ-type matrices: The virtual class of the family of conjugacy classes of matrices that are conjugate to JJ-type matrices is (q1)(q-1), namely that is given by the unique eigenvalue. In other words, [J/GL2(k)]=q1[J/\textup{GL}_{2}(k)]=q-1.

  • \blacksquare

    The MM-type matrices: The virtual class of the family of conjugacy classes of matrices that are conjugate to a diagonal matrix of two different eigenvalues can be computed as follows. We parametrize the family using the characteristic polynomial. The variety of all characteristic polynomials is isomorphic to 𝔸1×(𝔸1{0})\mathbb{A}^{1}\times(\mathbb{A}^{1}\setminus\{0\}), parametrizing the two coefficients (the constant term cannot be 0). The MM-type matrices correspond to characteristic polynomials with two different roots in this case. The scalar type matrices correspond to the characteristic polynomials that have a double root, meaning that the virtual class of the family of conjugacy classes of MM-type matrices is q(q1)(q1)=(q1)2q(q-1)-(q-1)=(q-1)^{2}.

So, the virtual class of all conjugacy classes of elements of GL2(k)\textup{GL}_{2}(k) is q1+q1+(q1)2=q21q-1+q-1+(q-1)^{2}=q^{2}-1.

Now, we turn our attention to the variety of commuting pairs, C2(GLr(k))C_{2}(\textup{GL}_{r}(k)). We begin by describing the pieces of C2(GLr(k))C_{2}(\textup{GL}_{r}(k)) as quotients by some subgroups of the symmetric group SrS_{r}. Consider a family of conjugacy classes given by a fixed type of Jordan normal form, 𝒥\mathcal{J}. In other words, we consider the subvariety C2𝒥(GLr(k))C_{2}^{\mathcal{J}}(\textup{GL}_{r}(k)) consisting of those commuting pairs (A,B)(A,B) where AA is conjugate to a matrix with a Jordan normal form of type 𝒥\mathcal{J}. The possible such Jordan normal forms form a variety that is the quotient of the space of possible eigenvalues modulo a group action that permutes the possible eigenvalues. We denote this quotient as E𝒥/H𝒥E_{\mathcal{J}}/H_{\mathcal{J}} where E𝒥E_{\mathcal{J}} denotes the space of possible eigenvalues and H𝒥H_{\mathcal{J}} denotes the group acting on it. Note that the group H𝒥H_{\mathcal{J}} is a product of symmetric groups, and thus it has a ”good” set of conjugacy classes of subgroups.

Using this, we get an algebraic map π:C2𝒥(GLr(k))E𝒥/H𝒥\pi:C_{2}^{\mathcal{J}}(\textup{GL}_{r}(k))\to E_{\mathcal{J}}/H_{\mathcal{J}} assigning to a pair (A,B)(A,B) of commuting matrices the conjugacy class of AA.

For example, consider the Jordan normal form

(λ1000λ0000μ1000μ)\left(\begin{array}[]{cccc}\lambda&1&0&0\\ 0&\lambda&0&0\\ 0&0&\mu&1\\ 0&0&0&\mu\\ \end{array}\right)

with λμ\lambda\neq\mu, then the corresponding variety is given as the quotient of (𝔸1{0})×(𝔸1{0})Δ𝔸1{0}(\mathbb{A}^{1}\setminus\{0\})\times(\mathbb{A}^{1}\setminus\{0\})\setminus\Delta_{\mathbb{A}^{1}\setminus\{0\}} by the group /2\mathbb{Z}/2\mathbb{Z} swapping the two coordinates (corresponding to the values of λ\lambda and μ\mu).

Consider the Cartesian product

X𝒥{{X_{\mathcal{J}}}}C2𝒥(GLr(k)){C_{2}^{\mathcal{J}}(\textup{GL}_{r}(k))}E𝒥{E_{\mathcal{J}}}E𝒥/H𝒥{E_{\mathcal{J}}/H_{\mathcal{J}}}π\scriptstyle{\pi}

where the bottom horizontal map is the quotient map E𝒥E𝒥/H𝒥E_{\mathcal{J}}\to E_{\mathcal{J}}/H_{\mathcal{J}}. The Cartesian product X𝒥X_{\mathcal{J}} is the variety of pairs (A,B)(A,B) where AA is conjugate to a matrix with Jordan type 𝒥\mathcal{J} and AA and BB commute. The representatives of the Jordan normal forms of the same type have the same centralizers (denoted by Z𝒥Z_{\mathcal{J}}), and therefore, X𝒥X_{\mathcal{J}} is isomorphic to E𝒥×(GLr(k)×Z𝒥/Z𝒥)E_{\mathcal{J}}\times(\textup{GL}_{r}(k)\times Z_{\mathcal{J}}/Z_{\mathcal{J}}) where the action of Z𝒥Z_{\mathcal{J}} on GLr(k)×Z𝒥\textup{GL}_{r}(k)\times Z_{\mathcal{J}} is given as

t.(g,z)=(gt,t1zg).t.(g,z)=(gt,t^{-1}zg).

Indeed, the map

E𝒥×(GLr(k)×Z𝒥)X𝒥E_{\mathcal{J}}\times(\textup{GL}_{r}(k)\times Z_{\mathcal{J}})\to X_{\mathcal{J}}

given by

(A,g,z)(gAg1,gzg1)(A,g,z)\mapsto(gAg^{-1},gzg^{-1})

is surjective, and the elements (A,g,z)(A,g,z) are mapped to the same element as the elements t.(A,g,z)=(A,gt,t1zt)t.(A,g,z)=(A,gt,t^{-1}zt) for tZ𝒥t\in Z_{\mathcal{J}}.

The action of H𝒥H_{\mathcal{J}} on E𝒥E_{\mathcal{J}} lifts to an action on X𝒥X_{\mathcal{J}}, and thus C2𝒥(GLr(k))C_{2}^{\mathcal{J}}(\textup{GL}_{r}(k)) is the quotient of E𝒥×(GLr(k)×Z𝒥/Z𝒥)E_{\mathcal{J}}\times(\textup{GL}_{r}(k)\times Z_{\mathcal{J}}/Z_{\mathcal{J}}) with the group H𝒥H_{\mathcal{J}} where H𝒥H_{\mathcal{J}} acts as

h.(A,g,z):=(PAP1,gP,P1zP)h.(A,g,z):=(PAP^{-1},gP,P^{-1}zP)

where PP is a permutation matrix corresponding to the element hH𝒥h\in H_{\mathcal{J}}. Lifting the action to E𝒥×GLr(k)×Z𝒥E_{\mathcal{J}}\times\textup{GL}_{r}(k)\times Z_{\mathcal{J}}, we obtain a Cartesian diagram

E𝒥×GLr(k)×Z𝒥{E_{\mathcal{J}}\times\textup{GL}_{r}(k)\times Z_{\mathcal{J}}}(E𝒥×GLr(k)×Z𝒥)/H𝒥{(E_{\mathcal{J}}\times\textup{GL}_{r}(k)\times Z_{\mathcal{J}})/H_{\mathcal{J}}}E𝒥×(GLr(k)×Z𝒥/Z𝒥){E_{\mathcal{J}}\times(\textup{GL}_{r}(k)\times Z_{\mathcal{J}}/Z_{\mathcal{J}})}C2𝒥(GLr(k)){C_{2}^{\mathcal{J}}(\textup{GL}_{r}(k))}E𝒥{E_{\mathcal{J}}}E𝒥/H𝒥.{E_{\mathcal{J}}/H_{\mathcal{J}}.}π\scriptstyle{\pi}

The map E𝒥×GLr(k)×Z𝒥EJE_{\mathcal{J}}\times\textup{GL}_{r}(k)\times Z_{\mathcal{J}}\to E_{J} is a GLr(k)×Z𝒥\textup{GL}_{r}(k)\times Z_{\mathcal{J}}-torsor. Since, the group GLr(k)\textup{GL}_{r}(k) and Z𝒥Z_{\mathcal{J}} are special linear algebraic groups in the sense of [3] (Z𝒥Z_{\mathcal{J}} is an extension of a smooth unipotent group by products of general linear groups), therefore,

[E𝒥×GLr(k)×Z𝒥/H𝒥]=[E𝒥/H𝒥][GLr(k)][Z𝒥].[E_{\mathcal{J}}\times\textup{GL}_{r}(k)\times Z_{\mathcal{J}}/H_{\mathcal{J}}]=[E_{\mathcal{J}}/H_{\mathcal{J}}][\textup{GL}_{r}(k)][Z_{\mathcal{J}}].

Similarly, the map

E𝒥×GLr(k)×Z𝒥E𝒥×(GLr(k)×Z𝒥/Z𝒥)E_{\mathcal{J}}\times\textup{GL}_{r}(k)\times Z_{\mathcal{J}}\to E_{\mathcal{J}}\times(\textup{GL}_{r}(k)\times Z_{\mathcal{J}}/Z_{\mathcal{J}})

is a Z𝒥Z_{\mathcal{J}}-torsor, thus

[E𝒥×GLr(k)×Z𝒥/H𝒥]=[C2𝒥(GLr(k))][Z𝒥].[E_{\mathcal{J}}\times\textup{GL}_{r}(k)\times Z_{\mathcal{J}}/H_{\mathcal{J}}]=[C_{2}^{\mathcal{J}}(\textup{GL}_{r}(k))][Z_{\mathcal{J}}].

This implies that [C2𝒥(GLr(k)]=[E𝒥/H𝒥][GLr(k)][C_{2}^{\mathcal{J}}(\textup{GL}_{r}(k)]=[E_{\mathcal{J}}/H_{\mathcal{J}}][\textup{GL}_{r}(k)] in K0(Stckk)\textup{K}_{0}(\textup{\bf{}Stck}_{k}).

Now, we are ready to prove Theorem 5.2.

Proof of Theorem 5.2.

The above implies that [C2(GLr(k))]=[GLr(k)]𝒥[E𝒥/H𝒥]=[GLr(k)][𝒞r][C_{2}(\textup{GL}_{r}(k))]=[\textup{GL}_{r}(k)]\sum_{\mathcal{J}}[E_{\mathcal{J}}/H_{\mathcal{J}}]=[\textup{GL}_{r}(k)][\mathcal{C}_{r}]. Using Theorem 5.3, we obtain

limr[C2(GLr(k))][GLr(k)]qr=1\lim_{r\to\infty}\frac{[C_{2}(\textup{GL}_{r}(k))]}{[\textup{GL}_{r}(k)]q^{r}}=1

in K0(Stckk)^\widehat{\textup{K}_{0}(\textup{\bf{}Stck}_{k})} proving our theorem. ∎

Remark 5.5.

In the proof of Theorem 5.2 we relied on two key statements: a) the virtual class of all conjugacy classes of GLr(k)\textup{GL}_{r}(k) is motivically stable, b) the groups GLr(k)\textup{GL}_{r}(k) and centralizer subgroups of the Jordan blocks of GLr(k)\textup{GL}_{r}(k) are special algebraic groups in the sense of [3]. These facts also hold for the family of SLr(k)\textup{SL}_{r}(k). In fact, a straight adaptation of the proof in [18] on the number of conjugacy classes of SLr(k)\textup{SL}_{r}(k) shows that the virtual class of the space of all conjugacy classes, [𝒞(SL2(k))][\mathcal{C}(\textup{SL}_{2}(k))] of SLr(k)\textup{SL}_{r}(k) is motivically stable with

limr[𝒞(SLr(k))]qr=1q1.\lim_{r\to\infty}\frac{[\mathcal{C}(\textup{SL}_{r}(k))]}{q^{r}}=\frac{1}{q-1}.

Thus, the following theorem can be deduced verbatim.

Theorem 5.6.

Consider the family of groups, SLr(k)\textup{SL}_{r}(k). Then,

limr[C2(SLr(k))]qr[SLr(k)]=1q1\lim_{r\to\infty}\frac{[C_{2}(\textup{SL}_{r}(k))]}{q^{r}[\textup{SL}_{r}(k)]}=\frac{1}{q-1}

in K0(Stckk)^\widehat{\textup{K}_{0}(\textup{\bf{}Stck}_{k})}.

References

  • [1] K. Behrend and A. Dhillon. On the motivic class of the stack of bundles. Advances in Mathematics, 212(2):617–644, 2007.
  • [2] N. Bridger and M. Kamgarpour. Character stacks are porc count. Journal of the Australian Mathematical Society, 115(3):289–310, 2023.
  • [3] C. Chevalley, A. Grothendieck, and J.-P. Serre. Anneaux de chow et applications, séminaire c. Chevalley. Chevalley, Ecole Normale Superieure, Paris, 1958.
  • [4] T. Church and B. Farb. Representation theory and homological stability. Advances in Mathematics, 245:250–314, 2013.
  • [5] T. Ekedahl. The Grothendieck group of algebraic stacks. arXiv e-prints, pages arXiv–0903, 2009.
  • [6] G. Frobenius. Über Gruppencharaktere. Sitzungsberichte der Königlich Preußischen Akademie der Wissenschaften zu Berlin, 1896.
  • [7] W. Fulton and J. Harris. Representation theory. Graduate Texts in Math, 129, 1991.
  • [8] S. M. Garge and A. Singh. Finiteness of z-classes in reductive groups. Journal of Algebra, 554:41–53, 2020.
  • [9] Á. González-Prieto. Topological Quantum Field Theories for Character Varieties. PhD thesis, Universidad Complutense de Madrid, 2018.
  • [10] Á. González-Prieto, M. Hablicsek, and J. Vogel. Arithmetic-geometric correspondence of character stacks via topological quantum field theory. arXiv preprint arXiv:2309.15331, 2023.
  • [11] Á. González-Prieto, M. Hablicsek, and J. Vogel. Virtual classes of character stacks. Journal of Geometry and Physics, page 105450, 2025.
  • [12] L. Göttsche. On the motive of the hilbert scheme of points on a surface. arXiv preprint math/0007043, 2000.
  • [13] M. Hablicsek and J. Vogel. Virtual classes of representation varieties of upper triangular matrices via topological quantum field theories. SIGMA. Symmetry, Integrability and Geometry: Methods and Applications, 18:095, 2022.
  • [14] N. Jacobson. Schur’s theorems on commutative matrices. Bulletin of the American Mathematical Society, 50(6):431–436, 1944.
  • [15] M. Kapranov. The elliptic curve in the s-duality theory and eisenstein series for kac-moody groups. arXiv preprint math/0001005, 2000.
  • [16] A. Kresch. Cycle groups for artin stacks. Inventiones Mathematicae, 138(3):495–536, 1999.
  • [17] S. Lang. Algebraic groups over finite fields. American Journal of Mathematics, 78(3):555–563, 1956.
  • [18] I. G. Macdonald. Numbers of conjugacy classes in some finite classical groups. Bulletin of the Australian Mathematical Society, 23(1):23–48, 1981.
  • [19] V. L. Popov and E. B. Vinberg. Invariant Theory, pages 123–278. Springer Berlin Heidelberg, Berlin, Heidelberg, 1994.
  • [20] D. A. Ramras and M. Stafa. Homological stability for spaces of commuting elements in lie groups. International Mathematics Research Notices, 2021(5):3927–4002, 2021.
  • [21] J. Schur. Zur theorie der vertauschbaren matrizen. 1905.
  • [22] U. Sharma and A. Singh. Commuting probability and simultaneous conjugacy classes of commuting tuples in a group. arXiv preprint arXiv:2002.01253, 2020.
  • [23] R. Vakil and M. M. Wood. Discriminants in the grothendieck ring. Duke Mathematical Journal, 164(6), 2015.
  • [24] J. Vogel. A topological quantum field theory for character varieties of non-orientable surfaces. arXiv preprint arXiv:2009.12310, 2020.
  • [25] J. Vogel. On the motivic higman conjecture. Journal of Algebra, 651:19–69, 2024.
  • [26] J. T. Vogel. Motivic invariants of character stacks. PhD thesis, PhD thesis. Universiteit Leiden, 2024. url: https://hdl. handle. net/1887 …, 2024.