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

Configuration spaces on a wedge of spheres and Hochschild–Pirashvili homology

Nir Gadish111Department of Mathematics, University of Michigan, 530 Church St, Ann Arbor, MI 48109, USA    Louis Hainaut222Department of Mathematics, Stockholm University, 106 91 Stockholm, Sweden
Abstract

We study the compactly supported rational cohomology of configuration spaces of points on wedges of spheres, equipped with natural actions of the symmetric group and the group Out(Fg)\operatorname{Out}(F_{g}) of outer automorphisms of the free group. These representations show up in seemingly unrelated parts of mathematics, from cohomology of moduli spaces of curves to polynomial functors on free groups and Hochschild–Pirashvili cohomology.

We show that these cohomology representations form a polynomial functor, and use various geometric models to compute many of its composition factors. We further compute the composition factors completely for all configurations of n10n\leq 10 particles. An application of this analysis is a new super-exponential lower bound on the symmetric group action on the weight 0 component of Hc(2,n)H^{*}_{c}(\mathcal{M}_{2,n}).

1 Introduction

This paper explores a circle of ideas centered around compactified configuration spaces of graphs and Hochschild cohomology, with applications to moduli spaces of marked curves and representation theory of outer automorphism groups of free groups. Revisiting ideas of Quoc Hô [Hô17], we explain the relation between Hochschild cohomology with square-zero coefficients and configuration spaces – two classes of mathematical objects that have independently been studied by many groups of researchers (see e.g. [TW19, PV18] for the former, and [Pet20, BCGY21] for the latter). This work seeks to popularize the connection, establish a dictionary, and bring the two points of view together to say new things about both objects as well as perform computations. Readers interested in either Hochschild homology, configuration spaces, polynomial representations of Out(Fn)\operatorname{Out}(F_{n}), moduli spaces of curves, or hairy graph complexes should be aware of how these subjects interact and share insights. The original motivation for this work is the following.

1.1 Super-exponential unstable cohomology of 2,n\mathcal{M}_{2,n}

The moduli space 2,n\mathcal{M}_{2,n} of genus 22 curves with nn marked points has complicated and mostly unknown rational cohomology. Considering the action of the symmetric group 𝔖n\mathfrak{S}_{n} by permuting marked points, the cohomology becomes an 𝔖n\mathfrak{S}_{n}-representation. In cohomological degree small compared to nn it exhibits representation stability as shown by Jiménez Rolland [Jim11], consequently bounding Betti numbers to be eventually polynomial in nn. In contrast, the orbifold Euler characteristic of 2,n\mathcal{M}_{2,n} was computed by Harer–Zagier to be (1)n+1240(n+1)!\frac{(-1)^{n+1}}{240}(n+1)! [HZ86, §6], thus necessitating super-exponential growth in nn as well as lots of odd-dimensional cohomology. This was awkward as there were almost no known infinite families of unstable cohomology classes, certainly not ones that grow super-exponentially in nn and reside in odd degrees (see the introduction of [BBP20] and Remark 1.3 in particular). Building on an unpublished formula of Petersen and Tommasi [PT] (see Proposition 1.10) we construct such families in the two highest non-trivial cohomological dimensions.

Our result is most naturally phrased Poincaré dually, on compact support cohomology. Below, F(X,n)F(X,n) is the ordered configuration space of nn distinct points in XX.

Theorem 1.1.

For every n3n\geq 3, the weight 0 rational cohomology gr0WHcn+(2,n;)\operatorname{gr}^{W}_{0}H^{n+*}_{c}(\mathcal{M}_{2,n};\mathbb{Q}) contains the following 𝔖n\mathfrak{S}_{n}-subrepresentations

{k odd sgnnH2(nk)(F(S3,n)/SU(2))k6Hn4(0,n)if =2k evensgnnH2(nk)(F(S3,n)/SU(2))k6if =3\begin{cases}\underset{k\text{ odd }}{\bigoplus}{\operatorname{sgn}_{n}\otimes H_{2(n-k)}(F(S^{3},n)/SU(2))}^{\oplus\left\lfloor\frac{k}{6}\right\rfloor}\oplus H_{n-4}(\mathcal{M}_{0,n})&\text{if }*=2\\ \underset{k\text{ even}}{\bigoplus}{\operatorname{sgn}_{n}\otimes H_{2(n-k)}(F(S^{3},n)/SU(2))}^{\oplus\left\lfloor\frac{k}{6}\right\rfloor}&\text{if }*=3\end{cases}

where x\lfloor x\rfloor is the integer part of xx. These degrees are Poincaré dual to the top two nontrivial degrees in which ordinary rational cohomology appears.

The characters of H(0,n)H_{*}(\mathcal{M}_{0,n}) and H(F(S3,n)/SU(2))H_{*}(F(S^{3},n)/SU(2)) have been computed by Getzler and Pagaria, respectively, and are recalled in §5.3.1. In particular, one can check that these subrepresentations have total dimension on the order of (n2)!(n-2)! thus grow super-exponentially.

Remark 1.2.

Note the rather peculiar appearance of configurations on odd dimensional manifolds. This brings to mind Hyde’s discovery [Hyd20] that the 𝔖n\mathfrak{S}_{n}-character of H(F(3,n))H^{*}(F(\mathbb{R}^{3},n)) appears in counts of polynomials over finite fields. Hyde’s formula was recently given a geometric explanation by Petersen–Tosteson [PT21]. An analogous geometric explanation of Corollary 1.1 would be most pleasing, but no such is known.

As explained in Proposition 1.10 below, there is a transferal of information from Hochschild homology to moduli spaces of curves. With this, calculations of Powell and Vespa of the former allow us to prove a conjecture from [BCGY21].

Corollary 1.3.

For all nn\in\mathbb{N}, the standard representation χ(n1,1)=stdn\chi_{(n-1,1)}=\operatorname{std_{n}} occurs in gr0WHcn+2(2,n)\operatorname{gr}^{W}_{0}H^{n+2}_{c}(\mathcal{M}_{2,n}) with multiplicity n12\left\lfloor\frac{n}{12}\right\rfloor, while the irreducible representation χ(2,1n2)=stdnsgnn\chi_{(2,1^{n-2})}=\operatorname{std_{n}}\otimes\operatorname{sgn}_{n} never occurs.

Our analysis suggests a multitude of new conjectures about the multiplicity of certain 𝔖n\mathfrak{S}_{n}-isotypical components inside gr0WHc(2,n)\operatorname{gr}_{0}^{W}H_{c}^{*}(\mathcal{M}_{2,n}), detailed in Conjecture 6.13

1.2 Configurations on wedges of spheres

We access the unstable cohomology of 2,n\mathcal{M}_{2,n} via configuration spaces of points on graphs. These are complicated topological spaces, which appear e.g. when studying moduli of tropical curves [BCGY21], geometric group theory and applied topology [GK02]. We take particular interest in their compactly supported cohomology, which turns out to depend only on the first Betti number of the graph (the loop order gg), and carries an action of the group Out(Fg)\operatorname{Out}(F_{g}) of outer automorphisms of the free group. These cohomology representations are of interest since, as we explain below, they are closely related to Hochschild–Pirashvili cohomology and play a universal role in the theory of ‘Outer’ polynomials functors on free groups. Notably, the cohomology provides a large natural class of Out(Fg)\operatorname{Out}(F_{g})-representations that do not factor through GLg()\operatorname{GL}_{g}(\mathbb{Z}). Detailed in §1.4, the special case of these representations with g=2g=2 is the key object that gives us a handle on the cohomology of 2,n\mathcal{M}_{2,n} and allows for the results mentioned above.

Recall the (ordered) configuration space of nn points on a topological space XX

F(X,n)={(x1,,xn)Xnxixjij}Xn.F(X,n)=\{(x_{1},\ldots,x_{n})\in X^{n}\mid x_{i}\neq x_{j}\;\forall i\neq j\}\subseteq X^{n}.

An often overlooked observation is that, while the homotopy type of the configuration space of n2n\geq 2 points is famously not a homotopy invariant of XX, the compactly supported cohomology Hc(F(X,n);)H^{*}_{c}(F(X,n);\mathbb{Z}) is an invariant of the proper homotopy type of XX under mild point-set hypotheses. We explain this in detail in §2.2. Let us consider cohomology with \mathbb{Q}-coefficients.

Since every finite graph is equivalent to a wedge of circles X=i=1gS1X=\bigvee_{i=1}^{g}S^{1}, it suffices to study such wedges. The group of homotopy classes of auto-equivalences of such XX is naturally isomorphic to Out(Fg)\operatorname{Out}(F_{g}), and thus Hc(F(i=1gS1,n);)H^{*}_{c}(F(\bigvee_{i=1}^{g}{S^{1}},n);\mathbb{Q}) acquires a natural Out(Fg)\operatorname{Out}(F_{g})-action.

Varied motivations, including the theory of string-links and spaces of embeddings mn\mathbb{R}^{m}\hookrightarrow\mathbb{R}^{n} [TW17, TW19], polynomial functors on free groups [HPV15, PV18], and unstable cohomology of moduli spaces, lead to the following hard open problem.

Goal 1.4.

Characterize the Out(Fg)\operatorname{Out}(F_{g})-representations Hc(F(i=1gS1,n);)H^{*}_{c}(F(\bigvee_{i=1}^{g}{S^{1}},n);\mathbb{Q}) for all finite wedges of circles.

This problem is essentially out of reach with our methods due to the fact that representations of Out(Fg)\operatorname{Out}(F_{g}) are in general not semi-simple. However, after semi-simplification they become a direct sum of factors arising from GLg()\operatorname{GL}_{g}(\mathbb{Z})-representations. The present work focuses therefore on the simpler question of determining only these composition factors, ignoring the extension problem of these factors.

We situate this problem as a special case of configurations on wedges of spheres of arbitrary dimension. Let X=iISdiX=\bigvee_{i\in I}S^{d_{i}} be a finite wedge of spheres, possibly of different dimensions di1d_{i}\geq 1. Our main object of study is the rational cohomology with compact support Hc(F(X,n);)H^{*}_{c}(F(X,n);\mathbb{Q}) (or rather an associated graded thereof), and how it behaves under continuous maps, in particular as the number and dimensions of spheres vary. This seemingly more general problem turns out to be no harder, as it is governed by a kind of polynomial functor which is entirely determined by its values on wedges of circles, as explained next.

Recall that a symmetric sequence is a sequence of vector spaces (Φ[m])m(\Phi[m])_{m\in\mathbb{N}} such that Φ[m]\Phi[m] carries an action by the symmetric group 𝔖m\mathfrak{S}_{m}333Such sequences are called linear species in combinatorics, and FB\operatorname{FB}-modules in the context of representation stability.. Joyal’s theory of analytic functors treats symmetric sequences as coefficients of a power series taking vector spaces to vector spaces:

VmΦ[m]𝔖mVm.V\longmapsto\bigoplus_{m\in\mathbb{N}}\Phi[m]\otimes_{\mathfrak{S}_{m}}V^{\otimes m}.

Such a functor is polynomial of degree at most dd if only the terms with mdm\leq d are nonzero. Our setup involves functors in two variables, hence we consider coefficients (Φ[n,m])(n,m)2(\Phi[n,m])_{(n,m)\in\mathbb{N}^{2}} which are graded 𝔖n×𝔖m\mathfrak{S}_{n}\times\mathfrak{S}_{m}-representations.

Theorem 1.5 (Polynomiality).

Consider the full subcategory of compact topological spaces XX that are homotopy equivalent to wedges of spheres, i.e. XiISdiX\simeq\bigvee_{i\in I}S^{d_{i}}. For every nn\in\mathbb{N} there exists a natural ‘collision’ filtration on the functor XHc(F(X,n))X\mapsto H_{c}^{\ast}(F(X,n)), whose associated graded factors through XH~(X)X\mapsto\tilde{H}^{*}(X) followed by a polynomial functor of degree nn.

Explicitly, there exists a collection of graded 𝔖n×𝔖m\mathfrak{S}_{n}\times\mathfrak{S}_{m}-representations Φ[n,m]\Phi[n,m], indexed by (n,m)2(n,m)\in\mathbb{N}^{2} and not depending on XX, and a natural 𝔖n\mathfrak{S}_{n}-equivariant isomorphism of graded vector spaces

grHc(F(X,n))mΦ[n,m]𝔖m(H~(X))m.\operatorname{gr}H_{c}^{*}(F(X,n))\cong\bigoplus_{m\in\mathbb{N}}{\Phi[n,m]\otimes_{\mathfrak{S}_{m}}(\tilde{H}^{\ast}(X))^{\otimes m}}. (1)

Here H~(X)iI[di]\tilde{H}^{\ast}(X)\cong\bigoplus_{i\in I}\mathbb{Q}[-d_{i}] is considered as a graded vector space, and \otimes is the graded tensor product whose symmetry uses the Koszul sign rule.

We will prove that Φ[n,m]=0\Phi[n,m]=0 whenever m>nm>n and that Φ[n,n]0\Phi[n,n]\neq 0, which means that the right hand side of (1) is the evaluation at H~(X)\tilde{H}^{*}(X) of a polynomial functor of degree nn.

Remark 1.6.

The associated graded is necessary here. For example when X=i=1gS1X=\bigvee_{i=1}^{g}{S^{1}} is a wedge of circles, both sides of (1) are representations of Out(Fg)\operatorname{Out}(F_{g}), but the right hand side always factors through a representation of GLg()\operatorname{GL}_{g}(\mathbb{Z}), while on the left hand side this is in general not the case without taking the associated graded – see Remark 4.11.

With this polynomiality theorem, understanding the ‘coefficient’ representations Φ[n,m]\Phi[n,m] is an important step towards the determination of Hochschild–Pirashvili cohomology for all wedges of spheres. Instead of Goal 1.4 we thus focus on the following. Below, we let χλ\chi_{\lambda} denote the irreducible representation of 𝔖n\mathfrak{S}_{n} associated with partition λn\lambda\vdash n.

Goal 1.7.

Compute Φ[n,m]\Phi[n,m] as a graded 𝔖n×𝔖m\mathfrak{S}_{n}\times\mathfrak{S}_{m}-representation. That is, compute the graded multiplicity of the 𝔖n×𝔖m\mathfrak{S}_{n}\times\mathfrak{S}_{m}-irreducible representation χλχμ\chi_{\lambda}\boxtimes\chi_{\mu} occurring in Φ[n,m]\Phi[n,m] for every pair of partitions λn\lambda\vdash n and μm\mu\vdash m.

This computation is in general a hard problem, though our geometric perspective gives access to important special cases presented in Theorem 1.8, and it further reveals structure that has not been explored until recently – e.g. the Lie structure discussed §3.3. We remark that Powell and Vespa study the same representations in [PV18] from a purely algebraic perspective; they also address there the problem of determining the extensions between the composition factors.

For XX a wedge of circles, Hc(F(X,n))H_{c}^{*}(F(X,n)) turns out to be concentrated in degrees =n1*=n-1 and nn only, see §4.3. Two classes of representations are particularly accessible.

Theorem 1.8 (Symmetric and exterior powers).

Let X=i=1gS1X=\bigvee^{g}_{i=1}S^{1} be a wedge of circles and let V=gH~(X)V=\mathbb{Q}^{g}\cong\tilde{H}^{\ast}(X) with its standard action by GLg()\operatorname{GL}_{g}(\mathbb{Z}). The 𝔖n\mathfrak{S}_{n}-equivariant isotypic component of the exterior power Λm(V)\Lambda^{m}(V) in the associated graded grHc(F(X,n))\operatorname{gr}H_{c}^{*}(F(X,n)) is given up to isomorphism by

grHcn(F(X,n))\displaystyle\operatorname{gr}H^{n}_{c}(F(X,n)) m0Hnm(0,n)Λm(H~(X))\displaystyle\geq\bigoplus_{m\geq 0}H_{n-m}(\mathcal{M}_{0,n})\otimes\Lambda^{m}(\tilde{H}^{*}(X)) (2)
grHcn1(F(X,n))\displaystyle\operatorname{gr}H^{n-1}_{c}(F(X,n)) m0Hnm2(0,n)Λm(H~(X))\displaystyle\geq\bigoplus_{m\geq 0}H_{n-m-2}(\mathcal{M}_{0,n})\otimes\Lambda^{m}(\tilde{H}^{*}(X)) (3)

For the symmetric power Symm(V)\operatorname{Sym}^{m}(V), its 𝔖n\mathfrak{S}_{n}-equivariant isotypic component in the associated graded grHc(F(X,n))\operatorname{gr}H_{c}^{*}(F(X,n)) is given by sign twists of the Whitehouse modules

grHcn(F(X,n))\displaystyle\operatorname{gr}H^{n}_{c}(F(X,n)) m0sgnnH2(nm)(F(3,n1))Symm(H~(X))\displaystyle\geq\bigoplus_{m\geq 0}\operatorname{sgn}_{n}\otimes H_{2(n-m)}(F(\mathbb{R}^{3},n-1))\otimes\operatorname{Sym}^{m}(\tilde{H}^{*}(X)) (4)
grHcn1(F(X,n))\displaystyle\operatorname{gr}H^{n-1}_{c}(F(X,n)) m0sgnnH2(nm1)(F(3,n1))Symm(H~(X))\displaystyle\geq\bigoplus_{m\geq 0}\operatorname{sgn}_{n}\otimes H_{2(n-m-1)}(F(\mathbb{R}^{3},n-1))\otimes\operatorname{Sym}^{m}(\tilde{H}^{*}(X)) (5)

with 𝔖n\mathfrak{S}_{n} acting via the identification F(3,n1)F(S3,n)/SU2F(\mathbb{R}^{3},n-1)\cong F(S^{3},n)/SU_{2}.

See §5.3 for proofs. The 𝔖n\mathfrak{S}_{n}-characters of H(0,n)H_{*}(\mathcal{M}_{0,n}) and H(F(3,n1))H_{*}(F(\mathbb{R}^{3},n-1)) are recalled in §5.3.1.

Encoding the ‘coefficient’ representations Φ[n,m]\Phi[n,m] by entries of a matrix indexed by partitions (λn,μm)(\lambda\vdash n,\mu\vdash m) as in Table 1 below, the last theorem fully describes the columns associated with partitions μ=(m)\mu=(m) and (1m)(1^{m}) for all m1m\geq 1.

Remark 1.9.

We note the unexpected appearance of the spaces 0,n\mathcal{M}_{0,n} and F(3,n1)F(\mathbb{R}^{3},n-1) while considering wedges of circles. This is a consequence of polynomiality in Theorem 1.5, and its uniform treatment of all spheres. In particular, for S2P1S^{2}\cong\mathbb{C}P^{1} and S3SU2S^{3}\cong SU_{2}.

1.3 Explicit computations

We approach calculation with two distinct tools, each giving access to one of the two symmetric group actions on Φ[n,m]\Phi[n,m] while obscuring the other.

  • A Chevalley–Eilenberg complex for Hc(F(X,n))H^{*}_{c}(F(X,n)), described by Petersen [Pet20], and an associated ‘collision’ spectral sequence. This gives insight into the right 𝔖m\mathfrak{S}_{m}-action on Φ[n,m]\Phi[n,m], determining the polynomial contribution of H~(X)\tilde{H}^{*}(X).

  • An 𝔖n\mathfrak{S}_{n}-equivariant cell structure on the one-point compactification F(X,n)+F(X,n)^{+}, when XX is a wedge of circles. This makes the 𝔖n\mathfrak{S}_{n}-action by permuting point labels relatively computable.

Overlaying the two sets of resulting data, we are able to compute many new irreducible multiplicities of Φ[n,m]\Phi[n,m]. In particualr, we calculate the 𝔖n×𝔖m\mathfrak{S}_{n}\times\mathfrak{S}_{m}-character of Φ[n,m]\Phi[n,m] completely for all n10n\leq 10, and a certain part of Φ[11,m]\Phi[11,m]. Beyond this range, the representations Φ[n,m]\Phi[n,m] are currently out of reach.

Our calculation technique is illustrated in Appendix A, culminating in Table 10 which shows all multiplicities for n=10n=10. Complete tabulations of Φ[n,m]\Phi[n,m] for all n10n\leq 10 can be found by following this URL444https://louishainaut.github.io/GH-ConfSpace/. All the computations for this paper were performed on Sage [Sage9.3]. Our algorithms are freely available on GitHub555https://github.com/louishainaut/GH-ConfSpace.

66 5,15,1 55 44 33 2,122,1^{2} 2,12,1 22 151^{5} 141^{4} 131^{3} 121^{2} 11
77 11 11 1
6,16,1
5,25,2 1 1 11 11
5,125,1^{2} 11 22 11 11
4,34,3 11 11 11
4,2,14,2,1 22 11 22 11 11 11
4,134,1^{3} 11 11 11
32,13^{2},1 11 22 11 11
3,223,2^{2} 22 11 11
3,2,123,2,1^{2} 11 11 22 11 11
3,143,1^{4} 11 11 11
23,12^{3},1 11 11 11
22,132^{2},1^{3} 11 11
2,152,1^{5}
171^{7} 11
Table 1: (λ,μ)(\lambda,\mu) multiplicity in Φ[7,m]\Phi[7,m] in codimension 1, arranged in lex-order of partitions. Unspecified entries are all zero. Yellow and red columns correspond to symmetric and alternating powers, respectively, and are completely described by Theorem 1.8. Lilac rows are computed in [PV18, Examples 3-4].

For example, Table 1 gives the complete decomposition of the lower degree of Φ[7,m]\Phi[7,m] into irreducibles for all mm. Equivalently, this is the irreducible decomposition of grHc6(F(gS1,7))\operatorname{gr}H^{6}_{c}(F(\bigvee_{g}S^{1},7)) as a representation of 𝔖7×Out(Fg)\mathfrak{S}_{7}\times\operatorname{Out}(F_{g}) up to splitting the collision filtration. Note that our Theorem 1.8 accounts for nearly all the nontrivial contributions. The cohomology in the other degree Hc7(F(gS1,7))H^{7}_{c}(F(\bigvee_{g}S^{1},7)) can be quickly obtained from Table 1 using the Euler characteristic.

The same Euler characteristic calculation gives a lower bound on the equivariant multiplicities of Hc(F(X,n))H^{*}_{c}(F(X,n)). It shows, in particular, that every Schur functor (see (9)) does occur in the cohomology, in fact super-exponentially many times as nn grows. See §5.4 and the example therein.

1.4 Weight 0 cohomology of moduli spaces

The coefficients Φ[n,]\Phi[n,-] discussed above determine the weight 0 cohomology groups gr0WHc(2,n,)\operatorname{gr}_{0}^{W}H_{c}^{*}(\mathcal{M}_{2,n},\mathbb{Q}), studied in recent work of Chan, Galatius and Payne [CGP22, CFGP19], via the following proposition. For a partition λm\lambda\vdash m let Φ[n,λ]\Phi[n,\lambda] denote the 𝔖n\mathfrak{S}_{n}-equivariant multiplicity space of the 𝔖m\mathfrak{S}_{m}-irreducible χλ\chi_{\lambda} in Φ[n,m]\Phi[n,m], equivalently it is Φ[n,m]𝔖mχλ\Phi[n,m]\otimes_{\mathfrak{S}_{m}}\chi_{\lambda}. Recall that Φ[,]\Phi[-,-] is graded, and denote by Φ[,]i\Phi[-,-]^{i} the part in degree ii.

Proposition 1.10.

For every partition with at most 22 parts λ=(a,b)\lambda=(a,b) there exists a coefficient rλr_{\lambda}\in\mathbb{N} such that for each i0i\geq 0 and nn\in\mathbb{N} there is an 𝔖n\mathfrak{S}_{n}-equivariant isomorphism

gr0WHci(2,n)λ=(a,b)(Φ[n,λ]i3|λ|)rλ\operatorname{gr}_{0}^{W}H_{c}^{i}(\mathcal{M}_{2,n})\cong\bigoplus_{\lambda=(a,b)}\left(\Phi[n,{\lambda}^{\ast}]^{i-3-|\lambda|}\right)^{\oplus r_{\lambda}} (6)

where gr0W\operatorname{gr}^{W}_{0} is the weight 0 subspace of HcH^{*}_{c} in the sense of Hodge theory, and the sum runs over partitions λ\lambda with up to two parts and conjugate partition λ{\lambda}^{\ast}.

Explicitly, the coefficients are given by

r(a,b)={ab6+1if a2b21ab6if a2b200if a2br_{(a,b)}=\begin{cases}\left\lfloor\frac{a-b}{6}\right\rfloor+1&\text{if }a\equiv_{2}b\equiv_{2}1\\ \left\lfloor\frac{a-b}{6}\right\rfloor&\text{if }a\equiv_{2}b\equiv_{2}0\\ 0&\text{if }a\not\equiv_{2}b\end{cases} (7)

where x\lfloor x\rfloor is the integer part of xx\in\mathbb{Q} and x2yx\equiv_{2}y is equivalence modulo 22.

Our Theorem 1.1 is a direct corollary of this Proposition along with Theorem 1.8

Note that the direct sum is finite since Φ[n,λ]=0\Phi[n,\lambda]=0 if |λ|>n|\lambda|>n. Also, due to the grading of Φ[n,λ]\Phi[n,\lambda], Formula (6) produces cohomology only in degrees n+2n+2 and n+3n+3.

Remark 1.11.

Formula (6) was discovered by Petersen and Tommasi [PT], and they explained it to us in private communication. We present here an alternative proof of this formula building on [BCGY21, Theorem 1.2]. Our determination of the coefficients rλr_{\lambda} is new.

The direct analog of (6) does not hold in higher genus g>2g>2. Instead, Petersen and Tommasi have informed us that it generalizes to the moduli space g,n\mathcal{H}_{g,n} of hyperelliptic curves, thus the calculations of this paper applies to the weight 0 compactly supported cohomology of those.

Our computation of Φ[n,λ]\Phi[n,\lambda] for all n10n\leq 10 mentioned in §1.3 recovers the 𝔖n\mathfrak{S}_{n}-character of gr0WHc(2,n)\operatorname{gr}^{W}_{0}H^{*}_{c}(\mathcal{M}_{2,n}) as appearing in [BCGY21]. We furthermore computed all terms Φ[11,(a,b)]\Phi[11,{(a,b)}^{\ast}], appearing in (6), thus obtaining the character of gr0WHc(2,11)\operatorname{gr}^{W}_{0}H^{*}_{c}(\mathcal{M}_{2,11}). We do not have explicit calculations beyond that. Nevertheless, in §6.1 we give a new lower bound on the character of gr0WHc(2,n)\operatorname{gr}^{W}_{0}H^{*}_{c}(\mathcal{M}_{2,n}), which grows very rapidly, and for n11n\leq 11 it accounts for a large portion of the full weight 0 cohomology.

1.5 Hochschild cohomology and configuration spaces

The Hochschild–Pirashvili homology of a wedge of circles is an invariant of a commutative algebra, which turns out to be fundamental to a wide range of fields: from rational homotopy theory of spaces of ‘long embeddings’ mn\mathbb{R}^{m}\hookrightarrow\mathbb{R}^{n} as studied by Turchin–Willwacher [TW17], to the theory of ‘Outer’ polynomial functors on the category of free groups studied by Powell–Vespa [PV18].

We show in §3, these cohomology groups are related via Schur–Weyl duality to Hc(F(i=1gS1,n);)H^{*}_{c}(F(\bigvee_{i=1}^{g}S^{1},n);\mathbb{Q}). This relation is known to some experts, but it does not appear in the literature in a form that we can readily use. For the benefit of the reader we give the following identification. Let A(n)=[ϵ1,,ϵn]/(ϵiϵjij)A_{(n)}=\mathbb{Q}[\epsilon_{1},\ldots,\epsilon_{n}]/(\epsilon_{i}\epsilon_{j}\mid i\leq j) be the square-zero algebra on nn generators, concentrated in degree 0, equipped with an 𝔖n\mathfrak{S}_{n}-action permuting the generators and multigraded by the degree of each generator; and let A(n)A^{(n)} be its linear dual coalgebra.

Theorem 1.12 (Hochschild cohomology and configurations).

Let XX be a simplicial set with finitely many non-degenerate simplices, and geometric realization |X||X|. For every positive integer nn the cohomology Hc(F(|X|,n),)H^{*}_{c}(F(|X|,n),\mathbb{Q}) is isomorphic to the (1,,1)(1,\ldots,1)-multigraded component of Hochschild–Pirashvili cohomology of X+=X{}X_{+}=X\sqcup\{*\} with coefficients in the coalgebra A(n)A^{(n)}. This isomorphism respects the 𝔖n\mathfrak{S}_{n}-action on both sides, and is natural with respect to maps XXX\to X^{\prime}.

In particular, for X=i=1gS1X=\bigvee_{i=1}^{g}S^{1}, whose group homotopy auto-equivalences is naturally identified with Out(Fg)\operatorname{Out}(F_{g}), this specializes to an isomorphism that respects the natural actions of 𝔖n\mathfrak{S}_{n} and Out(Fg)\operatorname{Out}(F_{g}) on both objects, with Out(Fg)\operatorname{Out}(F_{g}) identified with the homotopy equivalences i=1gS1i=1gS1\bigvee_{i=1}^{g}S^{1}\to\bigvee_{i=1}^{g}S^{1}.

We think of this relationship as a geometric interpretation of the algebraic construction of Hochschild–Pirashvili homology in [Pir00], and we prove it in §3.2. This theorem reformulates a hard open problem of Hochschild–Pirashvili (co)homology [TW19, §2.5] as the following.

1.6 Relations to previous work

The representations Hc(F(X,n))H^{*}_{c}(F(X,n)) appear in previous work in the following forms. Turchin–Willwacher define in [TW19, §2.5] a class of Out(Fg)\operatorname{Out}(F_{g})-representations which they call bead representations. They show that these do not factor through GL(g,)\operatorname{GL}(g,\mathbb{Z}), and are the smallest known representations with this property. Turchin–Willwacher pose the (still open) problem of describing the representations, and explain that they play a role in the rational homotopy theory of higher codimensional analogues of string links. Our current work provides a geometric interpretation of these representations as Hc(F(i=1gS1,n))H^{*}_{c}(F(\vee_{i=1}^{g}S^{1},n)), and computing our coefficients Φ[n,m]\Phi[n,m] is equivalent to calculating the composition factors of the bead representations.

In later work Powell–Vespa [PV18] discovered that the linear duals to the bead representations are fundamental to the theory of polynomial functors on the category of finitely generated free groups. They establish vast machinery to study these objects and compute the representations in some cases that are relevant to our work here. To make use of their work, and to make it more accessible to a topologically minded reader, we include a dictionary between their terminology and ours in 4.4. That section will recall the construction and significance of bead representations, and their relation to our work.

Another precursor to our project is Hô’s study of factorization homology and its relations to configuration spaces [Hô20]. On the one hand rational Hochschild homology is known to be a special case of factorization homology, and on the other hand Hô showed that many variants of configuration spaces arise as factorization homology with suitable coefficients [Hô20, Proposition 5.1.9]. Essentially our Theorem 1.12 is a special case of his work, but for the benefit of the reader we include our own proof.

1.7 Acknowledgments

We are grateful to Greg Arone, Dan Petersen, Orsola Tommasi and Victor Turchin for helpful conversations related to this work. We further thank Andrea Bianchi, Guillaume Laplante-Anfossi, Geoffrey Powell and Christine Vespa for their feedback on a previous version of the paper. The second author benefited from the guidance of Dan Petersen as his Ph.D. supervisor. We thank ICERM and Brown University for generously providing the computing resources on which we ran some of our programs. N.G. is supported by NSF Grant No. DMS-1902762; L.H. is supported by ERC-2017-STG 759082.

2 Preliminaries

2.1 Polynomial functors on Vect\textrm{Vect}_{\mathbb{Q}}

Eilenberg and MacLane introduced in [EM54] the notion of polynomiality for functors between categories of modules, defined using the notion of cross-effect functors. In the present work we use functors VectVect\textrm{Vect}_{\mathbb{Q}}\to\textrm{Vect}_{\mathbb{Q}}, sending (finite-dimensional) rational vector spaces to rational vector spaces. In this specific case the category of polynomial functors is semi-simple and the general theory of polynomial functors admits a simpler presentation, as recalled below. For additional details we refer to Joyal [Joy86] and Macdonald [Mac95, Appendix I.A]. The end of this section includes a brief discussion of the notion of polynomiality for functors 𝒈𝒓𝒑Vect\bm{grp}\to\textrm{Vect}_{\mathbb{Q}}, from finitely generated free groups to rational vector spaces, as considered in [DV15].

Definition 2.1 (Symmetric sequences).

A symmetric sequence (of \mathbb{Q}-vector spaces) is a sequence of vector spaces (Am)m=0(A_{m})_{m=0}^{\infty} such that AmA_{m} is equipped with a linear action of the symmetric group 𝔖m\mathfrak{S}_{m}. The sequence is finitely supported if Am=0A_{m}=0 for all m0m\gg 0; denoted as a finite sequence (Am)m=0n(A_{m})_{m=0}^{n} for some n0n\geq 0.

Proposition 2.2 (Polynomial functors, [Mac95, Appendix I.A]).

Fix n0n\geq 0 and let (Am)m=0n(A_{m})_{m=0}^{n} be a finitely supported symmetric sequence. A functor Ψ:VectVect\Psi\colon\textrm{Vect}_{\mathbb{Q}}\to\textrm{Vect}_{\mathbb{Q}} of the form

Ψ(V)=m=0nAm𝔖mVm\Psi(V)=\bigoplus_{m=0}^{n}{A_{m}\otimes_{\mathfrak{S}_{m}}{V^{\otimes m}}} (8)

is polynomial of degree n\leq n in the sense of Eilenberg–MacLane. Moreover, every polynomial functor Ψ:VectVect\Psi:\textrm{Vect}_{\mathbb{Q}}\to\textrm{Vect}_{\mathbb{Q}} is isomorphic to one obtained by the above construction, for representations AmA_{m} determined uniquely by Ψ\Psi up to isomorphism.

In the above proposition, the representation AmA_{m} is called the mm-th coefficient of the polynomial functor Ψ\Psi. The largest mm for which Am0A_{m}\neq 0 is the degree of Ψ\Psi.

Definition 2.3.

Analogously, for a general symmetric sequence (Am)m=0(A_{m})_{m=0}^{\infty}, the functor

Ψ(V)=m=0Am𝔖mVm\Psi(V)=\bigoplus_{m=0}^{\infty}{A_{m}\otimes_{\mathfrak{S}_{m}}{V^{\otimes m}}}

is called analytic.

Famously, analytic functors VectVect\textrm{Vect}_{\mathbb{Q}}\to\textrm{Vect}_{\mathbb{Q}} contain exactly the data of their sequence of coefficients (see e.g. [Mac95, Appendix I.A]). Stated precisely,

Proposition 2.4.

The construction (A0,,Am,)Ψ(A_{0},\ldots,A_{m},\ldots)\mapsto\Psi defines an equivalence of categories between the category of symmetric sequences and the category of analytic functors. It further restricts to an equivalence between the subcategory of finitely supported symmetric sequences and the subcategory of polynomial functors.

The inverse construction is given by sending Ψ\Psi to (A0,A1,)(A_{0},A_{1},\ldots) where AnA_{n} is the ‘multi-linear part’

An=Ψ(n)(1,,1)A_{n}=\Psi(\mathbb{Q}^{n})^{(1,\ldots,1)}

on which diagonal (n×n)(n\times n)-matrices act with weight (1,,1)(1,\ldots,1).

Remark 2.5.

In [Mac95] Macdonald gives a slightly different definition of polynomial functors than Eilenberg–Maclane’s, but this difference is immaterial in our context. In light of the previous proposition, one may take functors of the form (8) as a definition of polynomial functors VectVect\textrm{Vect}_{\mathbb{Q}}\to\textrm{Vect}_{\mathbb{Q}}. However, when Vect\textrm{Vect}_{\mathbb{Q}} is replaced with a different category of modules, there do exist polynomial functors that are not characterized by a symmetric sequence of coefficients. In that case, functors as in (8) define a proper subcategory of polynomial functors.

Given two polynomial functors Ψ1\Psi_{1} and Ψ2\Psi_{2} of respective degrees degΨ1=d1\deg\Psi_{1}=d_{1} and degΨ2=d2\deg\Psi_{2}=d_{2}, their sum ΦΨ\Phi\oplus\Psi, product ΦΨ\Phi\otimes\Psi, and composition ΦΨ\Phi\circ\Psi are also polynomial functors. Moreover

deg(Ψ1Ψ2)=max(d1,d2)deg(Ψ1Ψ2)=d1+d2deg(Ψ1Ψ2)=d1d2.\deg(\Psi_{1}\oplus\Psi_{2})=\max(d_{1},d_{2})\quad\deg(\Psi_{1}\otimes\Psi_{2})=d_{1}+d_{2}\quad\deg(\Psi_{1}\circ\Psi_{2})=d_{1}d_{2}.

Proposition 2.4 implies that polynomial functors form a semi-simple abelian category, and the irreducible objects are the so called Schur functors

𝕊λ:Vχλ𝔖mVm\mathbb{S}_{\lambda}:V\longmapsto\chi_{\lambda}\otimes_{\mathfrak{S}_{m}}V^{\otimes m} (9)

where χλ\chi_{\lambda} is the irreducible 𝔖m\mathfrak{S}_{m}-representation corresponding to the partition λm\lambda\vdash m. Most familiar are the symmetric and alternating powers 𝕊(m)(V)=Symm(V)\mathbb{S}_{(m)}(V)=\operatorname{Sym}^{m}(V) and 𝕊(1m)(V)=Λm(V)\mathbb{S}_{(1^{m})}(V)=\Lambda^{m}(V). Recall that the space 𝕊λ(V)\mathbb{S}_{\lambda}(V) is non-trivial if and only if the number of parts in λ\lambda does not exceed dimV\dim V, and in that case the resulting GL(V)\operatorname{GL}(V)-representations for different λ\lambda are all distinct.

More generally, if Ψ\Psi is a polynomial functor and its coefficients decompose as AmλmAλχλA_{m}\cong\bigoplus_{\lambda\vdash m}A_{\lambda}\otimes\chi_{\lambda}, i.e. the χλ\chi_{\lambda}-multiplicity space is AλA_{\lambda}, then

Ψ()λAλ𝕊λ()\Psi(-)\cong\bigoplus_{\lambda}A_{\lambda}\otimes\mathbb{S}_{\lambda}(-)

an irreducible decomposition of Ψ\Psi into Schur functors, with sum over all partitions λ\lambda.

In this paper we will actually not work with the category Vect\textrm{Vect}_{\mathbb{Q}} but with the category grVect\operatorname{grVect}_{\mathbb{Q}} of graded rational vector spaces, with the symmetry of \otimes obeying the Koszul sign rule; everything said until now applies verbatim to this situation. The coefficients of polynomial functors grVectgrVect\operatorname{grVect}_{\mathbb{Q}}\to\operatorname{grVect}_{\mathbb{Q}} can be determined by considering graded vector spaces VV of finite type, concentrated in a single degree:

Proposition 2.6.

Let V=d[i]V=\mathbb{Q}^{d}[-i] be a graded vector space of rank dd concentrated in degree ii, and let Ψ\Psi be a polynomial functor with coefficients AmλmAλχλA_{m}\cong\bigoplus_{\lambda\vdash m}{A_{\lambda}\otimes\chi_{\lambda}} where AλA_{\lambda} are graded vector spaces. Then

Ψ(V)={m0λmAλ𝕊λ(d)[mi]if i is evenm0λmAλ𝕊λ(d)[mi]if i is odd\Psi(V)=\begin{cases}\bigoplus_{m\geq 0}\bigoplus_{\lambda\vdash m}{A_{\lambda}\otimes\mathbb{S}_{\lambda}(\mathbb{Q}^{d})[-mi]}&\textrm{if $i$ is even}\\ \bigoplus_{m\geq 0}\bigoplus_{\lambda\vdash m}{A_{{\lambda}^{\ast}}\otimes\mathbb{S}_{\lambda}(\mathbb{Q}^{d})[-mi]}&\textrm{if $i$ is odd}\end{cases}

where the sum runs over all partitions and λ{\lambda}^{\ast} denotes the conjugate partition of λ\lambda.

In other words, when VV is concentrated in even degree, the value Ψ(V)\Psi(V) as a GL(V)\operatorname{GL}(V)-representation determines the coefficients of Ψ\Psi with at most dim(V)\dim(V) parts. On the other hand when VV is concentrated in odd degree, the GL(V)\operatorname{GL}(V)-representation Ψ(V)\Psi(V) determines the coefficients with all parts having size at most dim(V)\dim(V).

2.1.1 Polynomial functors from free groups

We also consider the notion of polynomiality for functors out of the category 𝒈𝒓𝒑\bm{grp} of finitely generated free groups. The relevant variant we need is that of contravariant functors Ψ:𝒈𝒓𝒑opVect\Psi\colon\bm{grp}^{op}\to\textrm{Vect}_{\mathbb{Q}}, however since this notion is not the main focus of this paper, we will only sketch the general idea and refer the interested reader to [HPV15, §3.2] for a detailed presentation.

The category 𝒈𝒓𝒑\bm{grp} consists of free groups Fda1,,adF_{d}\cong\langle a_{1},\ldots,a_{d}\rangle for every nonnegative integer dd and group homomorphisms. The free product FnFmFn+mF_{n}\ast F_{m}\cong F_{n+m} and the trivial group F0F_{0} endow this category with the structure of a pointed monoidal category; and [HPV15, §3] define contravariant polynomial functors of degree n1n-1 on 𝒈𝒓𝒑\bm{grp} to be ones for which the nn-th cross effect functor

crnΨ(Fg1,,Fgn):=ker(Ψ(Fg1Fgn)i=1nΨ(Fg1Fgi^Fgn))\operatorname{cr}_{n}\Psi(F_{g_{1}},\ldots,F_{g_{n}}):=\ker\left(\Psi(F_{g_{1}}*\cdots*F_{g_{n}})\to\bigoplus_{i=1}^{n}\Psi(F_{g_{1}}*\cdots*\widehat{F_{g_{i}}}*\cdots*F_{g_{n}})\right)

induced by the morphisms 1Fgi1\to F_{g_{i}} vanish identically.

Remark 2.7.

The reference [HPV15, §3], as well as a large part of the literature, focuses more on covariant polynomial functors. While the categories of covariant and contravariant polynomial functors on 𝒈𝒓𝒑\bm{grp} are not equivalent, the subcategories of functors taking finite-dimensional values are related via linear duality, see [Pow21, §7] for details. In particular, since the polynomial functors that we will encounter take finite-dimensional values, we will allow ourselves to dualize statements about covariant polynomial functors.

A major difference between polynomial functors 𝒈𝒓𝒑opVect\bm{grp}^{op}\to\textrm{Vect}_{\mathbb{Q}} and VectVect\textrm{Vect}_{\mathbb{Q}}\to\textrm{Vect}_{\mathbb{Q}} is that the former category is not semi-simple, and the analogue of Proposition 2.2 fails. Indeed, polynomial functors on 𝒈𝒓𝒑\bm{grp} have a more complicated structure, described next.

An important class of polynomial functors on 𝒈𝒓𝒑\bm{grp} is given by precomposing polynomial functors VectVect\textrm{Vect}_{\mathbb{Q}}\to\textrm{Vect}_{\mathbb{Q}} by the rationalization ι:𝒈𝒓𝒑Vect\iota\colon\bm{grp}\to\textrm{Vect}_{\mathbb{Q}}

Fd𝔞𝔟ddF_{d}\overset{\mathfrak{ab}}{\longmapsto}\mathbb{Z}^{d}\overset{\otimes\mathbb{Q}}{\longmapsto}\mathbb{Q}^{d}

and linear duality. In this way, every Schur functor 𝕊λ\mathbb{S}_{\lambda} as defined in (9) gives rise to a polynomial functor 𝒈𝒓𝒑opVect\bm{grp}^{op}\to\textrm{Vect}_{\mathbb{Q}} of degree |λ||\lambda|, denoted by ι𝕊λ\iota^{*}\mathbb{S}_{\lambda}. As before, these functors still comprise the simple polynomial functors, however now they admit non-trivial extensions.

Proposition 2.8 (Polynomial filtration, [DV15, Corollaires 3.6, 3.7]).

Every polynomial functor Ψ:𝐠𝐫𝐩opVect\Psi:\bm{grp}^{op}\to\textrm{Vect}_{\mathbb{Q}} admits a natural ‘polynomial’ filtration by degree. The associated graded functor splits as the direct sum of functors of the form ι𝕊λ\iota^{*}\mathbb{S}_{\lambda}.

Stated differently, every polynomial functor on 𝒈𝒓𝒑\bm{grp} is obtained as an iterated extension of functors ι𝕊λ\iota^{*}\mathbb{S}_{\lambda}. These functors ι𝕊λ\iota^{*}\mathbb{S}_{\lambda} are basic examples of the following notion.

Definition 2.9 (Outer functors).

A functor Ψ:𝒈𝒓𝒑opVect\Psi:\bm{grp}^{op}\to\textrm{Vect}_{\mathbb{Q}} is said to be an outer functor if it sends inner automorphisms to the identity map. That is, if for every free group FgF_{g} the Aut(Fg)\operatorname{Aut}(F_{g})-action on Ψ(Fg)\Psi(F_{g}) factors through Out(Fg)=Aut(Fg)/Inn(Fg)\operatorname{Out}(F_{g})=\operatorname{Aut}(F_{g})/\operatorname{Inn}(F_{g}) (recall that an automorphism is inner if it is the conjugation by a fixed element).

An outer functor is polynomial if it is both outer and polynomial in the sense discussed above.

2.2 Cohomology with compact support of configuration spaces

Homology and cohomology in this paper will be taken with rational coefficients unless otherwise stated, and we will henceforth suppress the coefficients from the notation. That is, Hc()H^{*}_{c}(-) will denote rational singular cohomology with compact support Hc(;)H^{*}_{c}(-;\mathbb{Q}).

Recall that Hc()H_{c}^{*}(-) is related to homology and cohomology as follows. First, when XX is a locally compact space then Hc(X)H~(X+)H_{c}^{*}(X)\cong\tilde{H}^{*}(X^{+}) for X+X^{+} the one-point compactification, and when XX is an nn-manifold Poincaré duality identifies Hci(X)Hni(X)Hni(X)H^{i}_{c}(X)\cong H_{n-i}(X)\cong H^{n-i}(X)^{\vee} (the second isomorphism needs the homology to be finite-dimensional). Furthermore, Hc(X)H^{*}_{c}(X) is itself the linear dual of Borel-Moore homology, which we denote by HBM(X)H^{BM}_{*}(X).

The configuration spaces F(X,n)F(X,n) are famously not a homotopy invariant of the space XX. For example the real line \mathbb{R} is homotopy equivalent to a point, but F({},2)=F(\{\ast\},2)=\emptyset while F(,2)F(\mathbb{R},2)\neq\emptyset; a highly nontrivial example of failure of homotopy invariance is found in [LS05]. Also, a continuous map f:XYf\colon X\to Y does not induce a map F(X,n)F(Y,n)F(X,n)\to F(Y,n) if ff is not injective and n2n\geq 2. We show however that at the level of cohomology with compact support homotopy invariance and functoriality do hold, at least for proper maps.

Proposition 2.10.

Let f:XYf\colon X\to Y be a proper map between locally compact Hausdorff spaces. For every positive integer nn, the map ff induces a natural morphism in each degree ii

Hci(F(Y,n))Hci(F(X,n)).H_{c}^{i}(F(Y,n))\to H_{c}^{i}(F(X,n)).
Proof.

Since ff is proper, the nn-fold cartesian product f×n:XnYnf^{\times n}\colon X^{n}\to Y^{n} is also proper. The subset U=(f×n)1(F(Y,n))U=(f^{\times n})^{-1}(F(Y,n)) is open in F(X,n)F(X,n) and the restriction of f×nf^{\times n} to UU is a proper map fn:=f×n|U:UF(Y,n)f_{n}:=f^{\times n}|_{U}\colon U\to F(Y,n). We can therefore construct the morphism Hci(F(Y,n))Hci(F(X,n))H_{c}^{i}(F(Y,n))\to H_{c}^{i}(F(X,n)) as the composition

Hci(F(Y,n))fnHci(U)Hci(F(X,n))H_{c}^{i}(F(Y,n))\overset{f_{n}^{*}}{\longrightarrow}H_{c}^{i}(U)\longrightarrow H_{c}^{i}(F(X,n)) (10)

where the second map is extension by zero.

Alternatively, ff induces a map F(X,n)+F(Y,n)+F(X,n)^{+}\to F(Y,n)^{+} between the one-point compactifications, where any collisions in YY are sent to \infty. The induced map on cohomology is, with the identification Hc(F(X,n))H~(F(X,n)+)H^{*}_{c}(F(X,n))\cong\tilde{H}^{*}(F(X,n)^{+}), the composition in (10). ∎

With this alternative definition it is easy to see that if f,g:XYf,g:X\to Y are homotopic through a proper homotopy H:X×IYH:X\times I\to Y, then they induce equal maps on Hc(F(,n))H^{*}_{c}(F(-,n)).

Corollary 2.11.

A proper homotopy equivalence f:XYf\colon X\to Y induces an isomorphism Hc(F(Y,n))Hc(F(X,n))H_{c}^{*}(F(Y,n))\cong H_{c}^{*}(F(X,n)) for each nn. In particular, if XX and YY are compact, the same follows for any homotopy equivalence.

A special case of this corollary is that for a finite connected graph GG, the groups Hc(F(G,n))H_{c}^{*}(F(G,n)) only depend on the first Betti number (or loop order) of GG.

Another implication of functoriality and homotopy invariance is the following.

Corollary 2.12.

Let XX be a compact space and nn a positive integer. The monoid Maps(X,X)\operatorname{Maps}({X},{X}) of continuous self maps acts on Hc(F(X,n))H_{c}^{\ast}(F(X,n)), and this action factors through its monoid of connected components [X,X]=π0(Maps(X,X))[X,X]=\pi_{0}(\operatorname{Maps}({X},{X})).

3 Geometry of Hochschild homology

This section highlights the geometric description of Hochschild–Pirashvili cohomology as compactly supported cohomology of a configuration space. This presentation highlights the fact that the cohomology acquires an additional Lie\operatorname{Lie}-module structure, discussed in §3.3. Readers interested only in our new calculations and results can safely skip this section.

Let us first briefly explain the notation used below. A more detailed introduction to Hochschild–Pirashvili cohomology can be found in [Pir00]. Throughout this section 𝕜\mathbbm{k} is any commutative ring.

Remark 3.1 (Basepoints).

If XX is a (simplicial) set, we will denote by X+X_{+} the pointed (simplicial) set obtained by adjoining to XX a disjoint basepoint. Let Fin\operatorname{Fin}_{\ast} be the category of finite pointed sets, a skeleton of which is given by the sets 𝐧+={,1,,n}\mathbf{n}_{+}=\{\ast,1,\ldots,n\} for all nn\in\mathbb{N}.

Let AA be an augmented commutative 𝕜\mathbbm{k}-algebra, with augmentation morphism ϵ:A𝕜\epsilon\colon A\to\mathbbm{k}. Define a covariant functor A:FinAlg𝕜A_{\bullet}\colon\operatorname{Fin}_{\ast}\to\operatorname{Alg}_{\mathbbm{k}}, sending 𝐧+\mathbf{n}_{+} to AnA^{\otimes n} and the morphism α:𝐦+𝐧+\alpha\colon\mathbf{m}_{+}\to\mathbf{n}_{+} to the morphism α:AmAn\alpha_{\ast}\colon A^{\otimes m}\to A^{\otimes n} defined on pure tensors by

α(a1am)=(j=1niα1(j)ai)ϵ(iα1()ai),\alpha_{\ast}(a_{1}\otimes\ldots\otimes a_{m})=\Big{(}\bigotimes_{j=1}^{n}{\prod_{i\in\alpha^{-1}(j)}a_{i}}\Big{)}\cdot\epsilon\Big{(}\prod_{i\in\alpha^{-1}(*)}{a_{i}}\Big{)},

with the convention that iai=1\prod_{i\in\emptyset}a_{i}=1 and a=1a_{*}=1. More conceptually, AA_{\bullet} is the unique coproduct-preserving functor FinAlg𝕜\operatorname{Fin}_{\ast}\to\operatorname{Alg}_{\mathbbm{k}} sending 𝟏+\mathbf{1}_{+} to AA and the map 𝟏+𝟎+\mathbf{1}_{+}\to\mathbf{0}_{+} to the augmentation map ϵ:A𝕜\epsilon\colon A\to\mathbbm{k}.

Definition 3.2 (Hochschild–Pirashvili homology).

Let X:ΔopFinX\colon\Delta^{op}\to\operatorname{Fin}_{\ast} be a pointed simplicial set with finitely many pp-simplices for all pp and let AA be an augmented commutative 𝕜\mathbbm{k}-algebra. The composition AXA_{\bullet}\circ X produces a simplicial 𝕜\mathbbm{k}-module, and its associated chain complex is denoted CH(X;A)CH_{\ast}(X;A). Define the Hochschild–Pirashvili homology of XX with coefficients in AA to be the homology of this chain complex, denoted HH(X;A)HH_{\ast}(X;A).

Remark 3.3.

The previous definition extends to general simplicial sets XX by considering the colimit of CH(X;A)CH_{\ast}(X^{\prime};A) over all finite simplicial subsets XXX^{\prime}\subseteq X, i.e. the left Kan extension. However, this extension will not be needed in our applications below.

When working instead with CC, a coaugmented cocommutative coalgebra, dualizing the previous definition gives HH(X;C)HH^{\ast}(X;C), the Hochschild–Pirashvili cohomology of XX with coefficients in CC. Explicitly, one defines a contravariant functor C:FinopMod𝕜C^{\bullet}:\operatorname{Fin}_{*}^{op}\to\operatorname{Mod}_{\mathbbm{k}} (in fact taking values in coalgebras), and the composition CXC^{\bullet}\circ X is a cosimplicial 𝕜\mathbbm{k}-module. The associated cochain complex CH(X;C)CH^{*}(X;C) has cohomology HH(X;C)HH^{\ast}(X;C) by definition.

Remark 3.4 (Loday construction).

Definition 3.2 is a special case of a more general construction, taking a pair (A,M)(A,M) of a commutative 𝕜\mathbbm{k}-algebra AA and a module MM over AA and producing a covariant functor (A,M):FinMod𝕜\mathcal{L}({A},{M})\colon\operatorname{Fin}_{*}\to\operatorname{Mod}_{\mathbbm{k}} (see [Pir00, §1.7]). Then HH(X;(A,M))HH_{\ast}(X;\mathcal{L}({A},{M})) is again defined as the homology of the chain complex associated to the simplicial 𝕜\mathbbm{k}-module (A,M)X\mathcal{L}({A},{M})\circ X. This construction has the property that HH(S1;(A,M))HH(A,M)HH_{\ast}(S^{1};\mathcal{L}({A},{M}))\cong HH_{\ast}(A,M), the classical Hochschild homology. The special case we consider in this paper is related to those in [TW19] and [PV18] via the following identification.

For XX a pointed simplicial set, AA an augmented 𝕜\mathbbm{k}-algebra and CC a coaugmented coalgebra, there exist isomorphisms (compare [PV18, Lemma 13.11])

CH(X+;A)=CH(X+;(A,𝕜))CH(X;(A,A)),CH_{\ast}(X_{+};A)=CH_{\ast}(X_{+};\mathcal{L}({A},{\mathbbm{k}}))\cong CH_{\ast}(X;\mathcal{L}({A},{A})), (11)
CH(X+;C)=CH(X+;(C,𝕜))CH(X;(C,C)),CH^{\ast}(X_{+};C)=CH^{\ast}(X_{+};\mathcal{L}({C},{\mathbbm{k}}))\cong CH^{\ast}(X;\mathcal{L}({C},{C})), (12)

natural in XX, AA and CC, and thus the Hochschild (co)homologies are the same.

To state the main claim of this section we introduce some terminology. Let VV be a free 𝕜\mathbbm{k}-module of finite rank, then the square-zero algebra AV:=𝕜VA_{V}:=\mathbbm{k}\oplus V is the unital 𝕜\mathbbm{k}-algebra with trivial multiplication on VV. Dually, define AV:=𝕜VA^{V}:=\mathbbm{k}\oplus V^{\vee} the coalgebra cogenerated by primitive elements VV^{\vee}, and note that AV(AV)A^{V}\cong(A_{V})^{\vee}. The central goal of the section is the following.

Proposition 3.5.

Let XX be a simplicial set with finitely many simplices in every degree, and for every n0n\geq 0 consider the coalgebra Anϵ1ϵnA^{\mathbb{Q}^{n}}\cong\mathbb{Q}\oplus\mathbb{Q}\epsilon_{1}\oplus\ldots\oplus\mathbb{Q}\epsilon_{n} cogenerated by the primitive cogenerators ϵ1,,ϵn\epsilon_{1},\ldots,\epsilon_{n}. There is an 𝔖n\mathfrak{S}_{n}-equivariant isomorphism

HH(X+;An)(1,,1)Hc(F(|X|,n))HH^{\ast}(X_{+};A^{\mathbb{Q}^{n}})^{(1,\ldots,1)}\cong H_{c}^{\ast}(F(|X|,n)) (13)

where the superscript (1,,1)(1,\ldots,1) denotes the multi-graded summand of Hochschild–Pirashvili cohomology that is multi-linear in the ϵi\epsilon_{i}’s (see Proposition 2.4).

Remark 3.6.

For a general simplicial set XX there is a homotopy invariant reformulation as

HH(X+;An)(1,,1)H~(|X|n/Δn(|X|)).HH^{\ast}(X_{+};A^{\mathbb{Q}^{n}})^{(1,\ldots,1)}\cong\tilde{H}^{\ast}(|X|^{n}/\Delta_{n}(|X|)).

This becomes expression (13) involving configuration spaces when |X||X| is compact.

Remark 3.7.

An equivalent formulation of Proposition 3.5 is the following: Consider the symmetric sequence given by nHc(F(|X|,n))n\mapsto H^{*}_{c}\left(F(|X|,n)\right). Then the corresponding functor

Vn0Hc(F(|X|,n))𝔖n(V)nV\mapsto\prod_{n\geq 0}H^{*}_{c}\left(F(|X|,n)\right)\otimes_{\mathfrak{S}_{n}}(V^{\vee})^{\otimes n} (14)

is naturally isomorphic to Hochschild–Pirashvili cohomology VHH(X+;AV)V\mapsto HH^{*}(X_{+};{A^{V}}). In particular the above functor is the linear dual of the functor at the subject of Powell–Vespa’s [PV18]. Hence the close relation between our work and theirs.

3.1 Labeled configuration spaces

We proceed by constructing a simplicial set whose simplicial homology with 𝕜\mathbbm{k}-coefficients is the Hochschild–Pirashvili homology. Let ComMon(Set)\operatorname{ComMon}(\operatorname{Set}_{\ast}) be the category of commutative monoids in the category of pointed sets, i.e. MComMon(Set)M\in\operatorname{ComMon}(\operatorname{Set}_{\ast}) is a pointed set equipped with a unit 1M1\in M and a product MMMM\wedge M\to M satisfying the usual axioms of commutative monoids (here \wedge is the smash product). Note that MM has a canonical augmentation ϵ:M{,1}\epsilon:M\to\{*,1\}, sending every non-invertible element to the basepoint.

Given a monoid MComMon(Set)M\in\operatorname{ComMon}(\operatorname{Set}_{\ast}), its Loday construction is defined analogously to Definition 3.2. This is the covariant functor M:FinComMon(Set)M_{\bullet}\colon\operatorname{Fin}_{\ast}\to\operatorname{ComMon}(\operatorname{Set}_{\ast}), defined on objects as 𝐧+Mn\mathbf{n}_{+}\mapsto M^{\wedge n}.

Definition 3.8 (Hochschild simplicial set).

Given MComMon(Set)M\in\operatorname{ComMon}(\operatorname{Set}_{\ast}) as above, and XX a pointed simplicial set with finitely many pp-simplices for every pp, the Hochschild simplicial set is defined as the composition MXM_{\bullet}\circ X.

As in Remark 3.3, this definition extends to general pointed XX by considering the colimit over finite simplicial subsets XXX^{\prime}\subseteq X.

Given a pointed set SS, let the reduced 𝕜\mathbbm{k}-module spanned by SS be 𝕜~[S]:=𝕜[S]/𝕜[]\tilde{\mathbbm{k}}[S]:=\mathbbm{k}[S]/\mathbbm{k}[*] where S*\in S is the basepoint. Then, 𝕜~[ST]𝕜~[S]𝕜~[T]\tilde{\mathbbm{k}}[S\vee T]\cong\tilde{\mathbbm{k}}[S]\oplus\tilde{\mathbbm{k}}[T] and 𝕜~[ST]𝕜~[S]𝕜~[T]\tilde{\mathbbm{k}}[S\wedge T]\cong\tilde{\mathbbm{k}}[S]\otimes\tilde{\mathbbm{k}}[T]. It is thus immediate that 𝕜~[]\tilde{\mathbbm{k}}[-] takes monoids in pointed sets to augmented 𝕜\mathbbm{k}-algebras. Furthermore, applying 𝕜~[]\tilde{\mathbbm{k}}[-] to a pointed simplicial set XX results in a simplicial 𝕜\mathbbm{k}-module, which under the Dold-Kan correspondence coincides with the reduced simplicial chain complex C~(X)\tilde{C}_{*}(X). These constructions are all compatible with \vee and \wedge of pointed simplicial sets.

Remark 3.9.

Due to the strong monoidality of 𝕜~[]\tilde{\mathbbm{k}}[-], it follows that the complex of reduced simplicial 𝕜\mathbbm{k}-chains of the Hochschild simplicial set MXM_{\bullet}\circ X is isomorphic to the Hochschild–Pirashvili chain complex CH(X;𝕜~[M])CH_{*}(X;\tilde{\mathbbm{k}}[M]).

In the rest of this subsection we introduce the notion of the configuration space of XX with labels in MM and prove that this simplicial set is isomorphic to the Hochschild simplicial set.

For every simplicial set XX, the Yoneda embedding provides a contravariant functor X:FinopsSetX^{\bullet}\colon\operatorname{Fin}_{\ast}^{op}\to\operatorname{sSet}_{\ast}, defined as the pointwise hom-functor Map(,X)\operatorname{Map}_{\ast}(-,X), sending 𝐧+\mathbf{n}_{+} to the pointed simplicial set Map(𝐧+,X)X×n\operatorname{Map}_{\ast}(\mathbf{n}_{+},X)\cong X^{\times n} whose basepoint is the constant pointed map.

Definition 3.10.

Given MComMon(Set)M\in\operatorname{ComMon}(\operatorname{Set}_{\ast}) and a pointed simplicial set XX, the configuration space of XX with labels in MM is the pointed simplicial set defined as the following coend

XFinM:=(n0X×nMn)/{X}^{\bullet}\otimes_{\operatorname{Fin}_{\ast}}{M}_{\bullet}:=\left.\left(\bigvee_{n\geq 0}X^{\times n}\wedge M^{\wedge n}\right)\middle/\sim\right. (15)

where \sim is the equivalence relation

(α(x1,,xn),(s1,,sm))((x1,,xn),α(s1,,sm))(\alpha^{\ast}(x_{1},\ldots,x_{n}),(s_{1},\ldots,s_{m}))\sim((x_{1},\ldots,x_{n}),\alpha_{\ast}(s_{1},\ldots,s_{m}))

for all pointed maps α:𝐦+𝐧+\alpha\colon\mathbf{m}_{+}\to\mathbf{n}_{+}.

Note in particular that if α\alpha is a bijection of 𝐧+\mathbf{n}_{+}, then α\alpha^{\ast} and α\alpha_{\ast} act on X×nX^{\times n} and MnM^{\wedge n} by mutually inverse permutations of the coefficients.

One can think of a point in the labeled configuration space as a tuple of points in XX, each of which decorated by an element of MM. When two such labeled points collide in XX, one replaces them by a single point labeled by the product of the two original labels. Points labeled by the unit 1M1\in M can be introduced or deleted freely.

Proposition 3.11.

Let XX be a pointed simplicial set and let MM be a commutative monoid in pointed sets. Then the Hochschild simplicial set MXM_{\bullet}\circ X is naturally isomorphic to the labeled configuration space XFinM{X}^{\bullet}\otimes_{\operatorname{Fin}_{\ast}}{M}_{\bullet}.

Proof.

When XX has finitely many nondegenerate simplices, this is essentially the co-Yoneda lemma enriched in simplicial sets. Indeed, at the level of pp-simplices the co-Yoneda lemma precisely gives a natural isomorphism

XpFinM=Map(,Xp)FinM=M(Xp){X_{p}}^{\bullet}\otimes_{\operatorname{Fin}_{\ast}}{M}_{\bullet}=\operatorname{Map}_{*}(-,X_{p})\otimes_{\operatorname{Fin}_{*}}M_{\bullet}=M_{\bullet}(X_{p})

and by naturality these isomorphisms form an isomorphism of simplicial sets.

When XX is any simplicial set, it is the filtered colimit over its finite simplicial subsets. Since maps from finite sets commute with filtered colimits, and Fin\otimes_{\operatorname{Fin}_{*}} commutes with all colimits, the same co-Yoneda argument extends to XX. ∎

Combining this proposition with Remark 3.9, one obtains the following.

Corollary 3.12.

Let XX be a pointed simplicial set and let MM be a commutative monoid in pointed sets. Then the Hochschild chain complex CH(X;𝕜~[M])CH_{*}(X;\tilde{\mathbbm{k}}[M]) is isomorphic to the reduced simplicial chain complex of the labeled configuration space XFinM{X}^{\bullet}\otimes_{\operatorname{Fin}_{\ast}}{M}_{\bullet} with 𝕜\mathbbm{k}-coefficients. Thus, the Hochschild–Pirashvili homology is the homology of the labeled configuration space.

The motivating example for this discussion is the subject of the next section.

3.2 Configuration spaces of distinct points

Let TT be a finite set. The square-zero monoid in pointed sets generated by TT is the pointed set M[T]:={0,1}TM[T]:=\{0,1\}\coprod T, with basepoint 0, unit 11 and otherwise trivial multiplication: ss=0s\cdot s^{\prime}=0 for all s,sSs,s^{\prime}\in S.

Claim 3.13.

Let X+X_{+} be a simplicial set with a disjoint basepoint and finitely many non-degenerate simplices. Given a finite set TT, the configuration space with labels in the square-zero monoid M[T]M[T] has geometric realization

|M[T](X+)|n0(F(|X|,n)×𝔖nTn)+|M[T]_{\bullet}\circ(X_{+})|\cong\bigvee_{n\geq 0}\left(F(|X|,n)\times_{\mathfrak{S}_{n}}T^{n}\right)^{+}

where Y+Y^{+} is the one-point compactification of a topological space YY. That is, the resulting labeled configuration space is the wedge sum of compactified configuration spaces of particles in |X||X| with decorations in TT.

Proof.

First, we prove that there is an isomorphism of simplicial sets

M[T](X+)n0(Xn/Δn(X))𝔖n(Tn)+M[T]_{\bullet}\circ(X_{+})\cong\bigvee_{n\geq 0}{\Big{(}X^{n}/{\Delta_{n}(X)}\Big{)}\wedge_{\mathfrak{S}_{n}}(T^{n})_{+}}

where Δn(X)Xn\Delta_{n}(X)\subseteq X^{n} is the fat diagonal – the simplicial subset whose pp-simplices are the nn-tuples of elements in XpX_{p} not all of whose coordinates are distinct; equivalently its pp-simplices are the non-injective functions from 𝒏\bm{n} to XpX_{p}.

On the left-hand side, pp-simplices are described as follows. There is a natural bijection M[T]((Xp)+)(({1}T)Xp)+M[T]_{\bullet}((X_{p})_{+})\cong\left((\{1\}\coprod T)^{X_{p}}\right)_{+}, further simplified by

({1}T)XpYXpTYn0Inj(𝒏,Xp)×𝔖nTn,\big{(}\{1\}\coprod T\big{)}^{X_{p}}\cong\coprod_{Y\subseteq X_{p}}T^{Y}\cong\coprod_{n\geq 0}\operatorname{Inj}(\bm{n},X_{p})\times_{\mathfrak{S}_{n}}T^{n}, (16)

where Inj(𝒏,Xp)\operatorname{Inj}(\bm{n},X_{p}) denotes the set of injective functions from 𝒏\bm{n} to XpX_{p}. Remembering the basepoint, we obtain the natural bijection

M[T]((Xp)+)n0Inj(𝒏,Xp)+𝔖n(Tn)+n0(Xpn/Δn(Xp))𝔖n(Tn)+,M[T]_{\bullet}((X_{p})_{+})\cong\bigvee_{n\geq 0}\operatorname{Inj}(\bm{n},X_{p})_{+}\wedge_{\mathfrak{S}_{n}}(T^{n})_{+}\cong\bigvee_{n\geq 0}\Big{(}X_{p}^{n}/{\Delta_{n}(X_{p})}\Big{)}\wedge_{\mathfrak{S}_{n}}(T^{n})_{+},

where the last isomorphism uses the obvious identification Xpn/Δn(Xp)Inj(𝒏,Xp)+X_{p}^{n}/\Delta_{n}(X_{p})\cong\operatorname{Inj}(\bm{n},X_{p})_{+} which sends noninjective functions in XpnX_{p}^{n} to the basepoint. It remains to verify that the simplicial maps on both sides commute with these bijections. Crucially, this uses the fact that the multiplication on TT is trivial.

Face and degeneracy maps α:XpXq\alpha:X_{p}\to X_{q} act on ({1}T)Xp(\{1\}\coprod T)^{X_{p}} by multiplying the labels of points in every fiber (with the empty product being 11). But since the product of two elements from TT is trivial, a function φ({1}T)Xp\varphi\in(\{1\}\coprod T)^{X_{p}} is sent to the basepoint unless every fiber of α\alpha contains at most one element with label in TT and every other element is labeled by the unit 11, in which case it is sent to the composition

α(φ1(T))α1φ1(T)𝜑T\alpha(\varphi^{-1}(T))\overset{\alpha^{-1}}{\longrightarrow}\varphi^{-1}(T)\overset{\varphi}{\longrightarrow}T

extended by 11 when this is undefined. On the RHS of (16) the function φ:φ1(T)T\varphi:\varphi^{-1}(T)\to T represents the element ((x1,φ(x1)),,(xn,φ(xn)))((x_{1},\varphi(x_{1})),\ldots,(x_{n},\varphi(x_{n}))) for some enumeration of φ1(T)\varphi^{-1}(T). This tuple is sent under α\alpha to ((α(x1),φ(x1)),,(α(xn),φ(xn)))((\alpha(x_{1}),\varphi(x_{1})),\ldots,(\alpha(x_{n}),\varphi(x_{n}))), which is easily seen to be the tuple corresponding to α(φ)\alpha_{*}(\varphi) described above.

Since geometric realization commutes with wedge sums, Cartesian products and quotients, one obtains a homeomorphism

|M[T](X+)|n0(|X|n/Δn(|X|))𝔖n(Tn)+,|M[T]_{\bullet}\circ(X_{+})|\cong\bigvee_{n\geq 0}{\Big{(}|X|^{n}/\Delta_{n}(|X|)\Big{)}\wedge_{\mathfrak{S}_{n}}(T^{n})_{+}},

where one observes that Δn(|X|)=|Δn(X)||X|n\Delta_{n}(|X|)=|\Delta_{n}(X)|\subseteq|X|^{n} is the topological fat diagonal, in which at least two coordinates are equal.

Up to this point, the analysis applies to any simplicial set. When |X||X| is compact (equivalently, XX has finitely many non-degenerate simplices), |X|n/Δn(|X|)|X|^{n}/\Delta_{n}(|X|) is homeomorphic to the one-point compactification F(|X|,n)+F(|X|,n)^{+}, inducing the claimed homeomorphism. ∎

Every labeled configuration space is naturally filtered by number of points in a configuration. In the case of the last claim the filtration actually splits, and the resulting graded factors are the compactified configuration spaces with a fixed number of points. This grading is refined further to a multi-grading by listing the multiplicities of every label sTs\in T. E.g. if T={red,blue}T=\{\text{red},\text{blue}\}, then the component with multi-degree (nr,nb)(n_{r},n_{b}) is the configuration space of nrn_{r} red points and nbn_{b} blue points, all of which distinct, compactified by one point at \infty.

Corollary 3.14.

For T={1,,n}T=\{1,\ldots,n\}, the (1,,1)(1,\ldots,1) multi-degree component of the realisation |M[T](X+)||M[T]_{\bullet}\circ(X_{+})| is the one-point compactification of the ordinary ordered configuration space of nn distinct points in |X||X|. The natural 𝔖n\mathfrak{S}_{n}-action on TT by permutations agrees with the usual action on the configuration space by relabeling points.

Remark 3.15.

From this point on we will abuse notation and use XX to refer both to the simplicial set and its geometric realization. The notation F(X,n)+F(X,n)^{+} for compactified configuration spaces can be understood either as a topological space or as the simplicial set Xn/Δn(X)X^{n}/\Delta_{n}(X), so that |F(X,n)+|F(|X|,n)+|F(X,n)^{+}|\cong F(|X|,n)^{+} when XX has finitely many non-degenerate simplices. Since simplicial and singular homology agree, the ambiguity in notation is immaterial.

Note that linearizing M[T]M[T] results in an augmented square-zero 𝕜\mathbbm{k}-algebra in the usual sense: 𝕜~[M[T]]𝕜1𝕜[T]\tilde{\mathbbm{k}}[M[T]]\cong\mathbbm{k}\cdot 1\oplus\mathbbm{k}[T] with trivial multiplication on 𝕜[T]\mathbbm{k}[T], denoted by ATA_{T}. Consider the Hochschild chain complex with coefficient in this square-zero algebra,

Corollary 3.16.

Let XX be a simplicial set with finitely many non-degenerate simplices. For T={1,,n}T=\{1,\ldots,n\} and AT=𝕜𝕜[T]A_{T}=\mathbbm{k}\oplus\mathbbm{k}[T] the square-zero 𝕜\mathbbm{k}-algebra generated by TT, the Hochschild chain complex CH(X+,AT)CH_{\ast}(X_{+},A_{T}) is naturally multi-graded by the multiplicity of each sTs\in T, so that the (1,,1)(1,\ldots,1) multi-degree component is equivalent to the reduced chains on the one-point compactification

CH(X+,AT)(1,,1)C~(F(X,n)+;𝕜)CH_{*}(X_{+},A_{T})^{(1,\ldots,1)}\simeq\tilde{C}_{*}\left(F(X,n)^{+};\mathbbm{k}\right) (17)

compatibly with the 𝔖n\mathfrak{S}_{n}-action on both chain complexes.

Similarly, the (d1,,dn)(d_{1},\ldots,d_{n}) multi-degree component is naturally isomorphic to the reduced chains on the one-point compactification of the configuration space of di\sum d_{i} distinct points with exactly did_{i} many having label iTi\in T.

Taking homology on both sides gives a natural isomorphism

HH(X+;AS)(1,,1)H~(F(X,n)+;𝕜).HH_{\ast}(X_{+};A_{S})^{(1,\ldots,1)}\cong\tilde{H}_{\ast}(F(X,n)^{+};\mathbbm{k}). (18)
Corollary 3.17.

With the notation of the previous corollary, and with AT:=𝕜1𝕜[T]A^{T}:=\mathbbm{k}\cdot 1\oplus\mathbbm{k}[T]^{\vee} the coalgebra dual to the square-zero algebra ATA_{T}, there is a natural isomorphism

HH(X+;AT)(1,,1)Hc(F(X,n);𝕜).HH^{*}(X_{+};A^{T})^{(1,\ldots,1)}\cong H^{*}_{c}\left(F(X,n);\mathbbm{k}\right). (19)
Proof.

We dualize the quasi-isomorphism in (17). On the left-hand side, Hochschild chains dualize to cochains. On the right-hand side, one uses the identification between cohomology with compact support and reduced cohomology of the one-point compactification. ∎

Specializing to rational coefficients and in light of the equivalence between symmetric sequences and analytic functors, these isomorphisms can be restated as the following isomorphisms, natural in the vector space VV:

HH(X+;AV)\displaystyle HH_{\ast}(X_{+};A_{V}) \displaystyle\cong n0H~(F(X,n)+;)𝔖nVn\displaystyle\bigoplus_{n\geq 0}{\tilde{H}_{\ast}(F(X,n)^{+};\mathbb{Q})\otimes_{\mathfrak{S}_{n}}V^{\otimes n}} (20)
HH(X+;AV)\displaystyle HH^{\ast}(X_{+};A^{V}) \displaystyle\cong n0Hc(F(X,n);)𝔖n(V)n\displaystyle\prod_{n\geq 0}{H_{c}^{\ast}(F(X,n);\mathbb{Q})\otimes_{\mathfrak{S}_{n}}(V^{\vee})^{\otimes n}} (21)

These facts are the ones claimed in Proposition 3.5.

3.3 Lie structure

The interpretation of Hochschild–Pirashvili homology as related to configuration space reveals additional structure, as explained to us by Victor Turchin.

Proposition 3.18.

Let XX be a compact CW complex. Then the symmetric sequence Hc(F(X,))H^{*}_{c}(F(X,\bullet)) is naturally endowed with a right module structure over the suspended Lie operad ΣLie\Sigma\operatorname{Lie}, that is a map Hc(F(X,))ΣLieHc(F(X,))H^{*}_{c}(F(X,\bullet))\circ\Sigma\operatorname{Lie}\to H^{*}_{c}(F(X,\bullet)) satisfying the usual right-module axioms.

The operadic suspension ΣLie\Sigma\operatorname{Lie} is defined by ΣLie(n)sgnnLie(n)[1n]\Sigma\operatorname{Lie}(n)\cong\operatorname{sgn}_{n}\otimes\operatorname{Lie}(n)[1-n], so that an algebra over it is equivalent to a Lie algebra structure on the (de)suspension.

Remark 3.19.

Around the time a first version of this article was made public, Christine Vespa informed us that Geoffrey Powell recently studied (outer) polynomial functors of 𝒈𝒓𝒑\bm{grp} via right Lie-modules [Pow21, Pow22a, Pow22], and the corresponding right Lie structure is the same as described in the previous Proposition. Briefly, Powell uses Morita theory to construct an equivalence of categories between outer polynomial functors on free groups – i.e. compatible sequences of Out(Fg)\operatorname{Out}(F_{g})-representations – and representations of the PROP associated with Lie\operatorname{Lie} operad. The equivalence centers around the representations Hc(F(X,n))H^{*}_{c}(F(X,n)) for XX a wedge of circles. See also [Ves22, §2.4-2.5].

Unpacking the definitions, this structure attaches a map

mφ:Hc(F(X,n1))Hc+1(F(X,n))m^{\varphi}:H^{*}_{c}(F(X,n-1))\to H^{*+1}_{c}(F(X,n))

to every surjection φ:[n][n1]\varphi:[n]\twoheadrightarrow[n-1], that are compatible with the symmetric group actions in the obvious way and satisfy appropriate versions of anti-symmetry and Jacobi identity.

The existence of such a structure follows from Koszul duality of the commutative and Lie operads, and the fact that the Lie\operatorname{Lie}-module nHc(F(X,n))n\mapsto H^{*}_{c}(F(X,n)) is Koszul dual to the Com\operatorname{Com}-module nH(Xn)n\mapsto H^{*}(X^{n}) (see [AT14, Lemma 11.4]). But let us be more explicit – the operations mφm^{\varphi} on Hc(F(X,n))H^{*}_{c}(F(X,n)) are obtained as follows.

Let XX be a compact CW complex and denote the hypersurface

Hij={(x1,,xn)Xnxi=xj}XnH_{ij}=\{(x_{1},\ldots,x_{n})\in X^{n}\mid x_{i}=x_{j}\}\subseteq X^{n}

for each iji\neq j so that the fat diagonal Δn(X)=Hij\Delta_{n}(X)=\bigcup H_{ij}. Any surjection [n][n1][n]\twoheadrightarrow[n-1] that sends ii and jj to the same image gives an identification Xn1HijXnX^{n-1}\cong H_{ij}\subseteq X^{n} by pullback, and the image of the (n1)(n-1)-st fat diagonal Δn1(X)Xn1\Delta_{n-1}(X)\subset X^{n-1} can be identified as follows.

Decomposing the nn-th fat diagonal as a union of two closed subspaces

Δn(X)=HijC where C={k,l}{i,j}Hkl\Delta_{n}(X)=H_{ij}\cup C\;\text{ where }\;C=\bigcup_{\{k,l\}\neq\{i,j\}}H_{kl} (22)

the intersection HijCH_{ij}\cap C is precisely the image of Δn1(X)\Delta_{n-1}(X) in HijH_{ij}. In particular, there is an isomorphism in relative cohomology

Hc(F(X,n1))H(Xn1,Δn1(X))H(Hij,HijC).H^{*}_{c}(F(X,n-1))\cong H^{*}(X^{n-1},\Delta_{n-1}(X))\cong H^{*}(H_{ij},H_{ij}\cap C). (23)

Now by excision in cohomology

H(Hij,HijC)H(Δn(X),C),H^{*}(H_{ij},H_{ij}\cap C)\cong H^{*}(\Delta_{n}(X),C), (24)

so the long exact sequence of the triple (XnΔn(X)C)(X^{n}\supset\Delta_{n}(X)\supset C) gives a coboundary map

H(Δn(X),C)𝑑H+1(Xn,Δn(X))Hc+1(F(X,n)).H^{*}(\Delta_{n}(X),C)\overset{d}{\longrightarrow}H^{*+1}(X^{n},\Delta_{n}(X))\cong H^{*+1}_{c}(F(X,n)). (25)

The composition of the above maps gives an operation

Hc(F(X,n1))H(Xn1,Δn1(X))Hc+1(F(X,n)),H^{*}_{c}(F(X,n-1))\cong H^{*}(X^{n-1},\Delta_{n-1}(X))\to H^{*+1}_{c}(F(X,n)), (26)

well-defined for a given choice of surjection [n][n1][n]\twoheadrightarrow[n-1].

Under the identification of symmetric sequences and analytic functors (Φ[])Ψ(\Phi[\bullet])\leftrightarrow\Psi, a right ΣLie\Sigma\operatorname{Lie}-module structure on a symmetric sequence (Φ[])(\Phi[\bullet]) becomes the map of analytic functors

Ψ(FreeLie(V))Ψ(V)\Psi\left(\operatorname{FreeLie}(V)\right)\to\Psi(V) (27)

such that every Lie-bracket increases the grading by +1+1.

Corollary 3.20.

Let XX be a simplicial set with finitely many simplices in every degree. The Hochschild–Pirashvili cohomology functor with square-zero coefficients HH(X;A)HH^{*}(X;A^{\bullet}) is endowed with the following natural structure:

HH(X;AFreeLie(V))HH(X;AV)HH^{*}(X;A^{\operatorname{FreeLie}(V)})\to HH^{*}(X;A^{V}) (28)

graded such that if FreeLiek(V)\operatorname{FreeLie}^{k}(V) denotes the subspace of kk-fold nested brackets then

HHi(X;AFreeLiek(V))HHi+k(X;AV).HH^{i}(X;A^{\operatorname{FreeLie}^{k}(V)})\to HH^{i+k}(X;A^{V}). (29)

If one is interested in Hochschild homology instead of cohomology, one need only dualize this Lie\operatorname{Lie} structure. More explicitly, the Hochschild–Pirashvili homology with square-zero coefficients carries a natural (shifted) right coLie\operatorname{Lie}-comodule structure.

Remark 3.21 (Geometric interpretation).

Thinking of Hochschild–Pirashvili cohomology HH(X;AV)HH^{*}(X;A^{V}) as related to cohomology of configurations of points in |X||X| with labels in VV, the Lie-module structure above comes from the following geometric construction.

Suppose a configuration {x1,,xn}|X|\{x_{1},\ldots,x_{n}\}\subseteq|X| with labels FreeLie(V)\operatorname{FreeLie}(V) has the point xnx_{n} labeled by the bracket [v,w][v,w]. The retraction φ:[n+1][n]\varphi:[n+1]\twoheadrightarrow[n] with φ(n+1)=φ(n)=n\varphi(n+1)=\varphi(n)=n induces the map

mφ:Hc(F(X,n))Hc+1(F(X,n+1))m^{\varphi}:H^{*}_{c}(F(X,n))\to H^{*+1}_{c}(F(X,n+1))

effectively splitting the nn-th point in two, as described above. Label the points of this new configuration so that xnx_{n} has label vv, xn+1x_{n+1} has label ww, while all other points retain their original label. This process can be repeated until there are no points labeled with brackets. The corollary above shows that the composition of these operations is independent of the order in which they were performed, and produces a well-defined cohomology class on configurations with labels in VV.

4 Geometric approaches to computation

Now let XX be a finite wedge of spheres. Our approach to computations is centered around contrasting two distinct tools for computing Hc(F(X,n))H^{*}_{c}(F(X,n)), each providing complementary information. First we use the so called collision spectral sequence, obtained by filtering the cartesian power XnX^{n} by its diagonals. This tool gives a handle on the polynomial structure of Hc(F(X,n))H^{*}_{c}(F(X,n)), but makes the symmetric group action by permutations less accessible.

Our second main tool, applying only when XX is a wedge of circles, is the cellular chain complex of the one-point compactification F(X,n)+F(X,n)^{+}. This second tool gives a handle on the symmetric group action on Hc(F(X,n))H^{*}_{c}(F(X,n)), but obscures the polynomial structure somewhat.

Our approach lets us calculate an associated graded grHc(F(X,n))\operatorname{gr}H^{*}_{c}(F(X,n)) for all n10n\leq 10. For example, a complete tabulation of the composition factors of Hc9(F(X,10))H^{9}_{c}(F(X,10)) can be found in Table 10. The output of all our calculations can be found by following this URL666https://louishainaut.github.io/GH-ConfSpace/.

4.1 Collisions and the CE-complex

We begin by considering the collision spectral sequence. This sequence has been used to study configuration spaces since the 80’s, see e.g. [CT78], [Kri94], [Tot96]. More recently, Hô [Hô17] and Petersen [Pet20] recast the spectral sequence as the Chevalley–Eilenberg complex of a twisted Lie algebra. The same spectral sequence also appears in Powell–Vespa’s [PV18] as coming from the polynomial filtration of Hochschild–Pirashvili homology, referred to as the Hodge filtration (though this is not in the sense of Hodge theory).

The following construction first appeared in Getzler’s [Get99a]. Recall that a twisted dg Lie algebra is a symmetric sequence 𝔤=(𝔤(n))n\mathfrak{g}=(\mathfrak{g}(n))_{n\in\mathbb{N}} of chain complexes that is a Lie algebra object in the category of symmetric sequences, with \otimes given by Day convolution: there is a bracket

[,]:Ind𝔖n×𝔖m𝔖n+m𝔤(n)𝔤(m)𝔤(n+m)[-,-]:\operatorname{Ind}_{\mathfrak{S}_{n}\times\mathfrak{S}_{m}}^{\mathfrak{S}_{n+m}}{\mathfrak{g}(n)\otimes\mathfrak{g}(m)}\to\mathfrak{g}(n+m)

that is 𝔖n+m\mathfrak{S}_{n+m}-equivariant and satisfies appropriate versions of anti-symmetry and Jacobi identity. Equivalently, 𝔤\mathfrak{g} is equipped with a left module structure over the Lie operad with respect to the composition product Lie𝔤𝔤\operatorname{Lie}\circ\mathfrak{g}\to\mathfrak{g}.

To every such 𝔤\mathfrak{g} one associates the Chevalley–Eilenberg bi-complex

CkCE(𝔤):=Symk(𝔤[1]),C^{CE}_{-k}(\mathfrak{g}):=\operatorname{Sym}^{k}(\mathfrak{g}[1]),

where we view 𝔤[1]\mathfrak{g}[1] as a bigraded object, with horizontal grading 1-1 and vertical grading given by the internal cohomological grading of 𝔤\mathfrak{g}. The differentials of this bi-complex are dd and δ\delta, where dd is the extension of differential on 𝔤\mathfrak{g} to a signed derivation on tensors, and

δ(sg1sgk)=i<jϵijs[gi,gj]sg1sgi^sgj^sgk,\delta(sg_{1}\otimes\ldots\otimes sg_{k})=\sum_{i<j}{\epsilon_{ij}s[g_{i},g_{j}]\otimes sg_{1}\otimes\ldots\otimes\widehat{sg_{i}}\otimes\ldots\otimes\widehat{sg_{j}}\otimes\ldots\otimes sg_{k}},

with ϵij\epsilon_{ij} being the sign induced by the Koszul sign rule when moving sgisg_{i} and sgjsg_{j} to the front. The bigrading of 𝔤[1]\mathfrak{g}[1] induces a bigrading of the double complex CCE(𝔤)C^{CE}_{*}(\mathfrak{g}). The sign ϵij\epsilon_{ij} depends on both the horizontal and the vertical grading. Note that since we use cohomological grading while the CE complex computes homology, the horizontal grading is negative.

Remark 4.1.

Since 𝔤\mathfrak{g} is a symmetric sequence, the Chevalley–Eilenberg homology HCE(𝔤)H_{*}^{CE}(\mathfrak{g}), defined as the cohomology of the total complex associated to CCE(𝔤)C_{*}^{CE}(\mathfrak{g}), admits a further filtration by arity. Moreover both differentials dd and δ\delta preserve the arity filtration, therefore the CE complex naturally splits by arity. With this, we note that if 𝔤(0)=0\mathfrak{g}(0)=0 (as will be the case in our situation), then CkCE(𝔤)(n)0C_{-k}^{CE}(\mathfrak{g})(n)\neq 0 only for nk1-n\leq-k\leq-1.

Further recall that if AA is any commutative dg algebra, then A𝔤=(A𝔤(n))nA\otimes\mathfrak{g}=(A\otimes\mathfrak{g}(n))_{n\in\mathbb{N}} admits a twisted Lie bracket by extending [,][-,-] in an AA-bilinear manner.

The central input to our calculation is the following result.

Proposition 4.2 ([Pet20], Corollary 8.8).

Let XX be a paracompact and locally compact Hausdorff space, and let AA be a cdga model for the compactly supported cochains Cc(X;)C_{c}^{*}(X;\mathbb{Q}). Then there is a natural isomorphism of symmetric sequences in graded vector spaces

Hc(F(X,n),)HCE(ASLie)(n){H_{c}^{*}(F(X,n),\mathbb{Q})}\cong H_{*}^{CE}(A\otimes\operatorname{S}\operatorname{Lie})(n)

where Lie\operatorname{Lie} is the Lie operad with its tautological Lie structure coming from operad multiplication, and SLie\operatorname{S}\operatorname{Lie} is its suspension as a twisted Lie algebra. More explicitly,

SLie(n)sgnnLie(n)[n]\operatorname{S}\operatorname{Lie}(n)\cong\operatorname{sgn}_{n}\otimes\operatorname{Lie}(n)[-n]

where Lie(n)\operatorname{Lie}(n) is the multi-linear part of the free Lie algebra on (x1,,xn)(x_{1},\ldots,x_{n}), with |xi|=0|x_{i}|=0, and V[n]V[-n] is the chain complex with VV placed in degree nn.

When AA models the cochains on a formal space, as is the case for wedges of spheres, the following simplification holds. Recall that a double complex is naturally filtered by its columns 𝔉p(s,tCs,t)=sp,tCs,t\mathfrak{F}_{p}\left(\bigoplus_{s,t}C_{s,t}\right)=\bigoplus_{s\geq p,t}C_{s,t}; we consider the resulting spectral sequence next.

Let (C(X),d)(C(X),d) be a functorial commutative cochain model for Cc(X;)C^{*}_{c}(X;\mathbb{Q}), where XX is any paracompact and locally compact Hausdorff space. Then the canonical filtration by columns of the bicomplex CCE(C(X)SLie)C^{CE}_{*}(C(X)\otimes\operatorname{S}\operatorname{Lie}) gives rise to a functorial spectral sequence converging to the symmetric sequence nHc(F(X,n);)n\mapsto{H^{*}_{c}(F(X,n);\mathbb{Q})}.

Lemma 4.3 (Formal spaces, compare [TW19, Remark 4.6]).

When XX is a compact formal space, i.e. the cdgas (C(X),d)(C(X),d) and (H(X),0)(H^{*}(X),0) are quasi-isomorphic, the spectral sequence induced by filtering CCE(C(X)SLie)C^{CE}_{*}(C(X)\otimes\operatorname{S}\operatorname{Lie}) by columns collapses at its E2E^{2}-page, and EE2HCE(H(X)SLie)E^{\infty}\cong E^{2}\cong H^{CE}_{*}(H(X)\otimes\operatorname{S}\operatorname{Lie}) functorially in XX.

Note 4.4.

Even when XX is formal, there might exist nonformal maps XXX\to X, i.e. such that the map C(X)C(X)C(X)\to C(X) is not equivalent to the induced H(X)H(X)H(X)\to H(X). If not for those, one would simply replace the model (C(X),d)(C(X),d) by (H(X),0)(H(X),0) and conclude that the CE-bicomplex splits into the direct sum of its rows. However, nonformal maps are responsible for nontrivial extensions between those rows, as is already the case for X=S1S1S1X=S^{1}\vee S^{1}\vee S^{1} (see Remark 4.11).

Proof.

Filtering the Chevalley–Eilenberg double complex by columns gives a spectral sequence with E0E^{0} coinciding with CCE(C(X)SLie)C^{CE}_{*}(C(X)\otimes\operatorname{S}\operatorname{Lie}) but only with the vertical differentials of C(X)C(X). Since CCEC^{CE}_{*} involves tensor powers and we are working over \mathbb{Q}, the Künneth formula gives a natural isomorphism

E1CCE(H(C(X)SLie))CCE(H(X)SLie)E_{1}\cong C^{CE}_{*}(H(C(X)\otimes\operatorname{S}\operatorname{Lie}))\cong C^{CE}_{*}(H(X)\otimes\operatorname{S}\operatorname{Lie}) (30)

where the latter isomorphism follows from the fact that ()SLie(-)\otimes\operatorname{S}\operatorname{Lie} is exact.

We claim that the spectral sequence collapses at its E2E^{2}-page. Indeed, since XX is compact and formal, the cohomology H(X)=H(X;)H(X)=H^{*}(X;\mathbb{Q}) itself constitutes a commutative cochain model for Cc(X)C_{c}^{\ast}(X) with vanishing differential. With this model the Chevalley–Eilenberg bicomplex computing Hc(F(X,n))H^{*}_{c}(F(X,n)) is exactly the E1E^{1}-page considered in the previous paragraph. But since the vertical differentials of the latter bicomplex are zero, its spectral sequence collapses at E2E^{2}. The fact that the spectral sequences for C(X)C(X) and H(X)H(X) both compute the same cohomology then forces all higher differentials of the former spectral sequence to also vanish. ∎

Terminology 4.5 (Collision filtration).

The collision filtration on Hc(F(X,n))H^{*}_{c}(F(X,n)) is the filtration induced by the following shift of the filtration 𝔉\mathfrak{F} by columns of the CE-bicomplex. For every nn\in\mathbb{N} let the pp-th level of the collision filtration be the (pn)(p-n)-th filtration by columns

FpHci(F(X,n))𝔉pnHi+nCE(C(X)SLie)(n).F_{p}\,H_{c}^{i}(F(X,n))\cong\mathfrak{F}_{p-n}\,H_{i+n}^{CE}(C(X)\otimes\operatorname{S}\operatorname{Lie})(n).

The reasoning behind the name is that this filtration comes from filtering the pair (Xn,Δn(X))(X^{n},\Delta^{n}(X)) by the number of collisions in the fat diagonal (see [BG19, §3.7] for details in the dual setting of Borel–Moore homology, but see the next remark).

Remark 4.6.

Comparing with [BG19, Definition 3.7.1], we note that this older definition was incorrect. Instead of defining the filtration directly on the homology, one should define the filtration FkCBM(Confn(X))F_{k}C_{*}^{BM}(\operatorname{Conf}^{n}(X)) at the chain level in the same fashion as described there, and consider the induced filtration on homology,

FkHBM(Confn(X)):=Im[H(FkCBM(Confn(X)))HBM(Confn(X))].F_{k}H_{*}^{BM}(\operatorname{Conf}^{n}(X)):=\operatorname{Im}\big{[}H_{*}(F_{k}C_{*}^{BM}(\operatorname{Conf}^{n}(X)))\to H_{*}^{BM}(\operatorname{Conf}^{n}(X))\big{]}.
Corollary 4.7.

When XX is a compact formal space, e.g. a finite wedge of spheres, then grFHc(F(X))\operatorname{gr}^{F}H_{c}^{*}(F(X)) is naturally isomorphic to the CE-homology HCE(H(X)SLie)H^{CE}_{*}(H(X)\otimes\operatorname{S}\operatorname{Lie}).

Moreover, for any map XYX\to Y between compact formal spaces, the induced map grFHc(F(Y,n))grFHc(F(X,n))\operatorname{gr}^{F}H^{*}_{c}(F(Y,n))\to\operatorname{gr}^{F}H^{*}_{c}(F(X,n)) is computed from the natural map H(Y)H(X)H^{*}(Y)\to H^{*}(X) via its action on the CE-homology HCE(H()SLie)H^{CE}_{*}(H^{*}(-)\otimes\operatorname{S}\operatorname{Lie}).

In particular, the action of the monoid Maps(X,X)\operatorname{Maps}({X},{X}) on grFHc(F(X,n))\operatorname{gr}^{F}H^{*}_{c}(F(X,n)) factors through End(H(X))\operatorname{End}(H^{*}(X)).

Unpacking the definition of the CE-complex when XX is a wedge of equidimensional spheres and forgetting the 𝔖n\mathfrak{S}_{n}-action, the first page of the collision spectral sequence takes the following explicit form involving tensor powers.

Proposition 4.8.

When X=i=1gSdX=\bigvee_{i=1}^{g}S^{d} is a wedge of equidimensional spheres so that Hd(X)gH^{d}(X)\cong\mathbb{Q}^{g}, the E1E_{1}-page of the collision spectral sequence takes the form

E1p,q{Tk(Hd(X))(|s(n,np)|(npk)) if q=dk and p+kn0 otherwiseE_{1}^{p,q}\cong\begin{cases}T^{k}(H^{d}(X))^{\oplus\left(|s(n,n-p)|\cdot\binom{n-p}{k}\right)}&\text{ if $q=dk$ and $p+k\leq n$}\\ 0&\text{ otherwise}\end{cases} (31)

with Tk(V)=VkT^{k}(V)=V^{\otimes k}, and |s(n,np)||s(n,n-p)| denoting an unsigned Stirling number of the first kind (see e.g. [Sta11, §1.3]).

Each of these E1p,qE_{1}^{p,q}-terms admits an action by 𝔖n\mathfrak{S}_{n}, but we do not have an explicit description of the latter in closed form.

Proof sketch.

Recall that Chevalley–Eilenberg complex of a twisted Lie algebra 𝔤\mathfrak{g} has underlying symmetric sequence of vector spaces the composition product Com𝔤\operatorname{Com}\circ\mathfrak{g}, with Com\operatorname{Com} the commutative operad. When 𝔤=(Hd(X))SLie\mathfrak{g}=(\mathbb{Q}\oplus H^{d}(X))\otimes\operatorname{S}\operatorname{Lie}, this composition consists of tensors power of Hd(X)H^{d}(X) tensored with the composition ComSLie\operatorname{Com}\circ\operatorname{S}\operatorname{Lie}.

On the other hand, applying Petersen’s formula for X=2X=\mathbb{R}^{2} identifies ComSLie\operatorname{Com}\circ\operatorname{S}\operatorname{Lie} with the cohomology of F(2,n)F(\mathbb{R}^{2},n) – the configuration spaces of nn points in the plane. Their Betti numbers are expressed in terms of Stirling numbers – see [Get99]. Keeping track of the filtration and tensor degree gives the stated formula. ∎

4.2 Action by outer automorphisms of free groups

The wedge i=1gS1\vee_{i=1}^{g}S^{1} is a classifying space for the free group FgF_{g}, thus its self-maps up to homotopy are given by End(Fg)\operatorname{End}(F_{g}) up to conjugation. As explained below, it follows that the graded vector spaces Hc(F(i=1gS1,n))H^{*}_{c}(F(\vee_{i=1}^{g}S^{1},n)) are representations of this endomorphism monoid (the statements in this section apply equally well to cohomology with integer coefficients, but we will not need that).

Pursuing further naturality, we use the description in terms of relative cohomology Hc(F(X,n))H(Xn,Δn(X))H^{*}_{c}(F(X,n))\cong H^{*}(X^{n},\Delta^{n}(X)) for every compact Hausdorff space XX, where Δn(X)\Delta^{n}(X) is the fat diagonal. Since i=1gS1\vee_{i=1}^{g}S^{1} is homotopy equivalent to the standard simplicial classifying space BFgBF_{g}, and since the functors XXnX\mapsto X^{n} and XΔn(X)X\mapsto\Delta^{n}(X) are homotopy invariant, there is a homotopy equivalence of pairs

((i=1gS1)n,Δn(i=1gS1))~((BFg)n,Δn(BFg))((\vee_{i=1}^{g}S^{1})^{n},\Delta^{n}(\vee_{i=1}^{g}S^{1}))\;\tilde{\longrightarrow}\;((BF_{g})^{n},\Delta^{n}(BF_{g})) (32)

with the latter pair obviously functorial in FgF_{g}. More precisely, let 𝐠𝐫𝐩\mathbf{grp} be the category of finitely generated free groups. The following is clear.

Proposition 4.9.

The construction FgH((BFg)n,Δn(BFg))F_{g}\mapsto H^{*}((BF_{g})^{n},\Delta^{n}(BF_{g})) is a contravariant functor from 𝐠𝐫𝐩\mathbf{grp} to graded vector spaces, and for every g1g\geq 1 it is isomorphic to Hc(F(gS1,n))H^{*}_{c}(F(\vee_{g}S^{1},n)) as 𝔖n\mathfrak{S}_{n}-representations.

In particular, for every gg and n1n\geq 1 the vector spaces Hc(F(gS1,n))H^{*}_{c}(F(\vee_{g}S^{1},n)) are equipped with a natural action of the monoid End(Fg)\operatorname{End}(F_{g}).

Every pointed map of spaces f:i=1gS1i=1hS1f:\vee_{i=1}^{g}S^{1}\to\vee_{i=1}^{h}S^{1} is determined up to homotopy by the induced homomorphism π1(f):FgFh\pi_{1}(f):F_{g}\to F_{h}, and so the above functor from 𝐠𝐫\mathbf{gr} completely characterizes the functoriality of Hc(F(i=1gS1,n))H^{*}_{c}(F(\vee_{i=1}^{g}S^{1},n)) on wedges of circles.

Of particular interest, every Hc(F(i=1gS1,n))H^{*}_{c}(F(\vee_{i=1}^{g}S^{1},n)) is a representation of the automorphism group Aut(Fg)\operatorname{Aut}(F_{g}). But since an inner automorphism – conjugation by some σFg\sigma\in F_{g} – is given by a map that is nonpointed-homotopic to the identity, its action on Hc(F(i=1gS1,n))H^{*}_{c}(F(\vee_{i=1}^{g}S^{1},n)) must be trivial. In other words, it is an outer functor (see Definition 2.9). Note that the configuration space functor F(,n)F(-,n) admits a natural 𝔖n\mathfrak{S}_{n}-action, and thus 𝔖n\mathfrak{S}_{n} acts on the resulting outer functor by natural transformations.

Lastly, pass to the associated graded of the collision filtration. By Corollary 4.7 the action of a homomorphism FgFhF_{g}\to F_{h} on grFHc(F(,n))\operatorname{gr}^{F}H^{*}_{c}(F(-,n)) is determined by the action on cohomology hg\mathbb{Z}^{h}\to\mathbb{Z}^{g}. Making this precise (compare with [PV18, Theorems 6.9, 17.8]),

Proposition 4.10.

Under the collision filtration, the associated graded of the outer functor FgHc(F(i=1gS1,n);)F_{g}\mapsto H^{*}_{c}(F(\vee_{i=1}^{g}S^{1},n);\mathbb{Q}) is the restriction of a well-defined functor on the category of finitely-generated free abelian groups along the abelianization,

FgabggrHc(F(i=1gS1,n))F_{g}\;\overset{\operatorname{ab}}{\longmapsto}\;\mathbb{Z}^{g}\longmapsto\operatorname{gr}H^{*}_{c}(F(\vee_{i=1}^{g}S^{1},n)) (33)

In particular, for every gg and n1n\geq 1 the graded quotients grHc(F(i=1gS1,n))\operatorname{gr}H^{*}_{c}(F(\vee_{i=1}^{g}S^{1},n)) admit a natural action of the matrix ring End(g)\operatorname{End}(\mathbb{Z}^{g}).

Proof.

Every homomorphism f:ghf:\mathbb{Z}^{g}\to\mathbb{Z}^{h} is the abelianization of some f¯:FgFh\bar{f}:F_{g}\to F_{h}. Define f:grHc(F(i=1hS1,n))grHc(F(i=1gS1,n))f^{*}:\operatorname{gr}H^{*}_{c}(F(\vee_{i=1}^{h}S^{1},n))\to\operatorname{gr}H^{*}_{c}(F(\vee_{i=1}^{g}S^{1},n)) by f¯\bar{f}^{*}. Corollary 4.7 shows that this does not depend on the choice of f¯\bar{f}. ∎

We will see below that (33) is in fact a polynomial functor on 𝐠𝐫𝐩\mathbf{grp} and that the collision filtration coincides with the polynomial filtration. Our eventual goal is to work towards computing its composition factors (see §5).

Remark 4.11 (Extensions of the Out\operatorname{Out}-action).

Since rational polynomial representations of GLg()\operatorname{GL}_{g}(\mathbb{Z}) admit no non-trivial extensions, one might expect that the extension problem for the collision filtration is trivial and thus that Hc(F(i=1gS1,n);)H^{*}_{c}(F(\vee_{i=1}^{g}S^{1},n);\mathbb{Q}) is isomorphic to HCE(H(i=1gS1)SLie)H^{CE}_{*}(H^{*}(\vee_{i=1}^{g}S^{1})\otimes\operatorname{S}\operatorname{Lie}) as representations.

However, this is not the case: the Out(Fg)\operatorname{Out}(F_{g})-representation Hc(F(i=1gS1,n);)H^{*}_{c}(F(\vee_{i=1}^{g}S^{1},n);\mathbb{Q}) does not factor through GLg()\operatorname{GL}_{g}(\mathbb{Z}), and the collision filtration exhibits enormously complicated nontrivial extensions – see [TW19, §2.3] and [PV18, Theorem 13].

4.3 Equivariant CW-structure

We now bring in a completely orthogonal approach to computing Hc(F(X,n);)H^{*}_{c}(F(X,n);\mathbb{Q}) when XX is a wedge of circles: using an 𝔖n\mathfrak{S}_{n}-equivariant CW-structure on the one-point compactifiction F(X,n)+F(X,n)^{+}. This was introduced and featured as the central computational tool in the first author’s recent work [BCGY21].

The resulting cellular cochain complex consists of only two nontrivial chain groups, each of which is free as an 𝔖n\mathfrak{S}_{n}-representation. The linear dual of this 22-step complex also featured in Powell–Vespa [PV18], though the vast generality of their framework needs some unpacking to make it amenable to computations.

Proposition 4.12 ([BCGY21, Theorem 1.2]).

Let X=i=1gS1X=\bigvee_{i=1}^{g}S^{1} be a wedge of gg circles. The 𝔖n\mathfrak{S}_{n}-representation Hc(F(X,n))H^{*}_{c}(F(X,n)) is computed by a 22-step complex of free 𝔖n\mathfrak{S}_{n}-modules

0[𝔖n](n+g2g1)cohomological degree n1[𝔖n](n+g1g1)cohomological degree n0\ldots\to 0\to\underbrace{\mathbb{Q}[\mathfrak{S}_{n}]^{\oplus\binom{n+g-2}{g-1}}}_{\text{cohomological degree }n-1}\to\underbrace{\mathbb{Q}[\mathfrak{S}_{n}]^{\oplus\binom{n+g-1}{g-1}}}_{\text{cohomological degree }n}\to 0\to\ldots (34)

As in the previous section, this statement holds equally well for integral cohomology, but we will not need that here.

Proof.

First note that Hc(F(X,n))H^{*}_{c}(F(X,n)) coincides with the ordinary reduced cohomology of the one-point compactification F(X,n)+F(X,n)^{+}. Then [BCGY21] gives an 𝔖n\mathfrak{S}_{n}-equivariant CW-structure on this compactification, with cells in dimensions n1n-1 and nn only and a free 𝔖n\mathfrak{S}_{n}-action. The resulting cellular cochain complex is as claimed. ∎

[BCGY21] also gives explicit formulas for the coboundary map and for the actions of endomorphisms of XX on this complex. We do use these in explicit calculations, but their details are not important for our discussion.

What is important is that the complex splits into its isotypic components. That is, for every irreducible representation χλ\chi_{\lambda} of 𝔖n\mathfrak{S}_{n}, the multiplicity space χλ𝔖nHc(F(X,n))\chi_{\lambda}\otimes_{\mathfrak{S}_{n}}H^{*}_{c}(F(X,n)) is also computed by the same complex

0χλ(n+g2g1)χλ(n+g1g1)0\ldots\to 0\to\chi_{\lambda}^{\oplus\binom{n+g-2}{g-1}}\to\chi_{\lambda}^{\oplus\binom{n+g-1}{g-1}}\to 0\to\ldots (35)

where one only needs to specialize the coboundary map to χλ\chi_{\lambda}. This small complex efficiently computes the multiplicity of χλ\chi_{\lambda} in Hc(F(X,n))H^{*}_{c}(F(X,n)) for various values of nn and gg. Its downside, however, is that the collision filtration is not as readily accessible as in the Chevalley–Eilenberg complex of §4.1. The following discussion explains how to ‘see’ the collision filtration on the terms of (35).

Remark 4.13 (End(Fg)\operatorname{End}(F_{g})-action).

Recall that upon taking the associated graded of the collision filtration on Hc(F(X,n);)H^{*}_{c}(F(X,n);\mathbb{Q}), the End(Fg)\operatorname{End}(F_{g})-action discussed in §4.2 factors through the matrix ring End(g)\operatorname{End}(\mathbb{Z}^{g}) (see Corollary 4.10). Furthermore, on the associated graded, the filtration degree is exhibited by the weights of the action of the diagonal matrices gEnd(g)\mathbb{Z}^{g}\subset\operatorname{End}(\mathbb{Z}^{g}). However, these weights can already be read-off from simple chain-level endomorphisms of the 2-step complex (35).

Lemma 4.14.

Let X=i=1gS1X=\bigvee_{i=1}^{g}S^{1} be a wedge of gg circles and let χλ\chi_{\lambda} be an irreducible representation of 𝔖n\mathfrak{S}_{n}. The action of the diagonal matrix diag(d1,,dg)Mg()\operatorname{diag}(d_{1},\ldots,d_{g})\in M_{g}(\mathbb{Z}) on the χλ\chi_{\lambda}-multiplicity space of grHc(F(X,n);)\operatorname{gr}H^{*}_{c}(F(X,n);\mathbb{Q}) is realized by the (non-invertible) space-level map φ:XX\varphi:X\to X that for every 1ig1\leq i\leq g wraps the ii-th circle around itself with degree did_{i}.

This induces an operator on the 2-step complex (35) that preserves each χλ\chi_{\lambda}-summand, and is thus an effectively computable block-diagonal transformation whose eigenspaces determine the collision filtration.

Proof.

The transformation φ\varphi acts on the homology of XX by the diagonal matrix diag(d1,,dg)\operatorname{diag}(d_{1},\ldots,d_{g}). Since the induced map on F(X,n)+F(X,n)^{+} is equivariant and cellular, it further induces an endomorphism of the complex (35).

The result of wrapping the ii-th circle around itself is realized as a sum of shuffles of the points lying on that circle. But regardless of their order, the number of points on each circle is preserved by the operation. These numbers of points are the invariants that differentiate the 𝔖n\mathfrak{S}_{n}-orbits of cells in F(X,n)+F(X,n)^{+}, hence the summands of (35) are preserved. In particular, the chain operator φ\varphi_{*} can be diagonalized within every summand.

Lastly, since a scalar matrix diag(n,,n)\operatorname{diag}(n,\ldots,n) acts on H1(X)H^{1}(X) by the scalar nn, will act on the subspace of collision filtration degree p\leq p by eigenvalues ndn^{d} with npdnn-p\leq d\leq n. In this way the eigenspaces of φ\varphi detect the collision filtration at the chain level. ∎

Remark 4.15.

Varying the diagonal entries (d1,,dg)(d_{1},\ldots,d_{g}) in the above lemma, the trace by which φ\varphi acts on (35) characterizes the GLg()\operatorname{GL}_{g}(\mathbb{Z})-action on grHc(F(X,n))\operatorname{gr}H^{*}_{c}(F(X,n)) completely, which is the ultimage goal of this project – see §5 below.

Furthermore, the block-diagonal structure of the action of φ\varphi on (35) is effective for computer calculations: it can be diagonalized on every χλ\chi_{\lambda} summand individually, thus allowing for parallel calculations on relatively small matrices.

4.4 Bead representations

Let us recall where the 𝔖n×Out(Fg)\mathfrak{S}_{n}\times\operatorname{Out}(F_{g})-representations Hc(F(i=1gS1,n))H^{*}_{c}(F(\bigvee_{i=1}^{g}S^{1},n)) previously appeared in the literature and establish a dictionary with related work. Turchin and Willwacher studied the Hochschild–Pirashvili cohomology of i=1gS1\bigvee_{i=1}^{g}S^{1}, and in [TW19, Section 2.5] they consider the (1,,1)(1,\ldots,1)-multigraded component, equipped with its 𝔖n\mathfrak{S}_{n}-action – by Proposition 3.5 this is the same as our Hc(F(i=1gS1,n))H^{*}_{c}(F(\bigvee_{i=1}^{g}S^{1},n)). They then split the 𝔖n\mathfrak{S}_{n}-action into isotypic components and call the resulting Out(Fg)\operatorname{Out}(F_{g})-representation bead representations.

Proposition 4.16.

For every λn\lambda\vdash n, Turchin–Willwacher’s bead representations UλI,UλIIU_{\lambda}^{I},U_{\lambda}^{II} are the Out(Fg)\operatorname{Out}(F_{g})-equivariant multiplicity of the Specht module χλ\chi_{\lambda} in Hc(F(i=1gS1,n))H^{*}_{c}(F(\bigvee_{i=1}^{g}S^{1},n)) for =n*=n and n1n-1 respectively. That is,

Uλ(g)=Hc(F(i=1gS1,n))𝔖nχλ.U_{\lambda}^{*}(g)=H^{*}_{c}(F(\vee_{i=1}^{g}S^{1},n))\otimes_{\mathfrak{S}_{n}}\chi_{\lambda}. (36)

They pose a (still open) problem to describe these representations, starting with computing the decomposition of grHc(F(i=1gS1,n))\operatorname{gr}H^{*}_{c}(F(\vee_{i=1}^{g}S^{1},n)) into Schur functors. This latter point is exactly the subject of this paper, e.g. they are the rows of Table 1. The decomposition of all UλIU_{\lambda}^{I} and UλIIU_{\lambda}^{II} with |λ|=n10|\lambda|=n\leq 10 can be found on this webpage777https://louishainaut.github.io/GH-ConfSpace/.

Remark 4.17 (Extensions of Schur functors).

The bead representations turn out to be central to the theory of outer polynomial functors (see Definition 2.9). Indeed, recall that the irreducible polynomial GLg()\operatorname{GL}_{g}(\mathbb{Z})-representations are given by Schur functors 𝕊λ(g)\mathbb{S}_{\lambda}(\mathbb{Q}^{g}), which admit no non-trivial extensions. On the other hand, every Schur functor can be pulled back to Out(Fg)\operatorname{Out}(F_{g}), and as such they do admit extensions, i.e. there exist non-split surjections of Out(Fg)\operatorname{Out}(F_{g})-representations Eι𝕊λ(g)E\twoheadrightarrow\iota^{*}\mathbb{S}_{\lambda}(\mathbb{Q}^{g}). Amazingly, Powell–Vespa give a canonical surjection Hcn(F(X,n))𝔖nχλι𝕊λH^{n}_{c}(F(X,n))\otimes_{\mathfrak{S}_{n}}\chi_{{\lambda}^{\ast}}\twoheadrightarrow\iota^{*}\mathbb{S}_{\lambda} and prove that it is the maximal indecomposable extension of outer polynomial functors, i.e. a projective cover (see [PV18, Theorem 19.1] for the dual statement). Moreover, every outer polynomial functor admits a minimal projective resolution by sums of bead representations.

Decomposing the bead representations into their composition factors is thus a fundamental task in the representation theory of Out(Fg)\operatorname{Out}(F_{g}), serving as further motivation for our present study.

Powell and Vespa prove a multitude of facts about these functors in [PV18], and we think it valuable to make their results accessible to the topologically minded reader. We will therefore devote the rest of this section bridging the terminology gap between their setup and what we consider to be more natural in the topological context.

  • Powell–Vespa study the Hochschild–Pirashvili homology of XX that we discuss in §3. In (20) we highlight its relation to homology of compactified configuration spaces as a functor in both a space XX and a vector space VV. Powell–Vespa denote this bi-functor by (X,V)HH(X,(AV,AV))(X,V)\mapsto HH_{*}(X,\mathcal{L}({A_{V}},{A_{V}})), which is isomorphic to our HH(X+,AV)HH_{*}(X_{+},A_{V}) in (20). This is an analytic functor in VV, and for fixed XX our Proposition 3.5 identifies the corresponding symmetric sequence of coefficients as the Borel–Moore homology nHBM(F(X,n))n\mapsto H_{*}^{BM}(F(X,n)), linearly dual to Hc(F(X,n))H_{c}^{*}(F(X,n)). In Powell–Vespa’s notation, this symmetric sequence of coefficients is the functor HH(X;ϑInjFin)HH_{*}(X;\vartheta^{*}\operatorname{Inj}^{\textbf{Fin}}).

  • They get a functor from the category of free groups by composing with the classifying space B():FgB(Fg)B(-):F_{g}\mapsto B(F_{g}) in place of the space XX. Since B(Fg)B(F_{g}) is coherently homotopy equivalent to the wedge i=1gS1\vee_{i=1}^{g}S^{1}, their functors agree with the ones we study here.

  • They show that these Hochschild homology groups form an outer polynomial functor on the category of free groups, and so over a field of characteristic 0 all their composition factors are of the form ι𝕊λ\iota^{*}\mathbb{S}_{\lambda} for various partitions λ\lambda. They denote these functors by αSλ\alpha S_{\lambda}. Calculating their multiplicities is the subject of our work.

    The associated graded representation we call grHc(F(i=1gS1,n))\operatorname{gr}H_{c}^{*}(F(\vee_{i=1}^{g}S^{1},n)) corresponds in their formalism to αncrnHH(B();ϑInjFin)\alpha_{n}\operatorname{cr}_{n}HH_{*}(B(-);\vartheta^{*}\operatorname{Inj}^{\textbf{Fin}}), evaluated at FgF_{g}. Here crn\operatorname{cr}_{n} is the functor that extracts the nn-th coefficient of a polynomial functor, returning an 𝔖n\mathfrak{S}_{n}-representation, and αn\alpha_{n} converts this representation into a sum of Schur functors.

  • For an integer partition λn\lambda\vdash n, a subscript λ\lambda either on Hochschild homology HH(;ϑInjFin)HH_{*}(-;\vartheta^{*}\operatorname{Inj}^{\textbf{Fin}}) or on the coefficients ϑInjFin\vartheta^{*}\operatorname{Inj}^{\textbf{Fin}} refers to the χλ\chi_{\lambda}-multiplicity space of the corresponding 𝔖n\mathfrak{S}_{n}-representation.

    In particular, our coefficients Φ[λ,m]\Phi[\lambda,m] for a partition λ\lambda and m|λ|m\leq|\lambda| is the 𝔖m\mathfrak{S}_{m}-representation crmHH(B();ϑInjλFin)\operatorname{cr}_{m}HH_{*}(B(-);\vartheta^{*}\operatorname{Inj}^{\textbf{Fin}}_{\lambda}), with Φ[λ,μ]\Phi[\lambda,\mu] giving the multiplicity of αSμ\alpha S_{{\mu}^{\ast}} (note that the partition μ\mu needs to be transposed).

  • As they discuss in [PV18, §16.3], for λn\lambda\vdash n a partition of nn, the bead representations UλIU_{\lambda}^{I} and UλIIU_{\lambda}^{II}, mentioned at the beginning of §4.4, are dual to HHn(B();ϑInjλFin)HH_{n}(B(-);\vartheta^{*}\operatorname{Inj}^{\textbf{Fin}}_{\lambda}) and HHn1(B();ϑInjλFin)HH_{n-1}(B(-);\vartheta^{*}\operatorname{Inj}^{\textbf{Fin}}_{\lambda}) respectively. Powell–Vespa refer to these by ωβn(Sλ)\omega\beta_{n}(S_{\lambda}) and (Cokerad¯coalgΣ)λ(\operatorname{Coker}_{\overline{\operatorname{ad}}}\mathbb{P}^{\Sigma}_{\operatorname{coalg}})_{\lambda} respectively.

Example 4.18 ((2,1n2)(2,1^{n-2}) bead representation).

To illustrate the translation from Powell–Vespa’s formalism, consider [PV18, Example 4]:

HH(B();ϑInj(2,1n2)Fin){αS(n1,1)=n0otherwise,HH_{*}(B(-);\vartheta^{*}\operatorname{Inj}^{\textbf{Fin}}_{(2,1^{n-2})})\cong\begin{cases}\alpha S_{(n-1,1)}&*=n\\ 0&\text{otherwise},\end{cases} (37)

the =n*=n case is also denoted ωβnS(2,1n2)\omega\beta_{n}S_{(2,1^{n-2})}.

We reinterpret this line as stating that for XX a wedge of circles, the cohomology Hc(F(X,n))H^{*}_{c}(F(X,n)) has as its χ(2,1n2)\chi_{(2,1^{n-2})}-multiplicity space isomorphic to the Schur functor ι𝕊(n1,1)(H~1(X))\iota^{*}\mathbb{S}_{{(n-1,1)}^{\ast}}(\tilde{H}^{1}(X)) when =n*=n, and otherwise vanishes. In terms of the coefficients Φ[,]\Phi[-,-], this means that the graded vector space Φ[(2,1n2),(2,1n2)]\Phi[(2,1^{n-2}),(2,1^{n-2})] has rank 11 concentrated in degree 0, and that for every other partition λ(2,1n2)\lambda\neq(2,1^{n-2}) the graded vector space Φ[(2,1n2),λ]\Phi[(2,1^{n-2}),\lambda] is trivial.

For completeness of the dictionary, we include:

  • In their calculations Powell–Vespa frequently use the functors βdSλ\beta_{d}S_{\lambda}, which in our terminology are linear dual to the top cochains in the 22-step complex (35) equipped with Aut(Fg)\operatorname{Aut}(F_{g})-actions . These also assemble to a polynomial functor on free groups, but it does not factor through Out(Fg)\operatorname{Out}(F_{g}), i.e. conjugations act nontrivially as noted in [BCGY21, Remark 2.12].

5 Polynomiality and consequences

Let us next consider configurations on XX for XX a wedge of spheres. As we let the number of spheres vary, the compactly supported cohomology acquires the structure of a polynomial functor evaluated on the vector space H~(X)\tilde{H}^{*}(X), as shown in this section.

Remark 5.1.

Having a single polynomial functor compute the cohomology of F(X,n)F(X,n) for any finite wedge of spheres constrains the functor and endows it with further structure that one would not have expected. For example consider the following three cases:

  1. 1.

    For configurations on a wedge of 11-spheres, the cohomology is closely related to the algebraic construction of Hochschild–Pirashvili homology as an exponential functor, see [PV18]. On the other hand, the configuration spaces admit a Fox-Neuwirth cell decomposition888This is a decomposition into locally closed sets that become cells of the one-point compactification. with cells in only the top two dimensions, freely permuted by the symmetric group. This gives a free presentation of the cohomology as an 𝔖n\mathfrak{S}_{n}-module, leading to rather efficient calculations – details in §4.3.

  2. 2.

    For configurations on a wedge of 22-spheres, the ambient space X=i=1gS2X=\bigvee_{i=1}^{g}S^{2} can be realized as a complex algebraic curve of genus gg. The cohomology in question thus admits a mixed Hodge structure, and an action by the Galois group of \mathbb{Q}. Moreover, for a single sphere F(P1,n)PSL2()×0,nF(\mathbb{C}P^{1},n)\cong PSL_{2}(\mathbb{C})\times\mathcal{M}_{0,n}, which explains the appearance of these moduli spaces in our calculations below.

  3. 3.

    For configurations on a wedge of 33-spheres, since S3SU(2)S^{3}\cong SU(2) is a Lie group, it follows that F(S3,n)S3×F(3,n1)F(S^{3},n)\cong S^{3}\times F(\mathbb{R}^{3},n-1). This lets us identify the equivariant multiplicity of all exterior powers.

5.1 Polynomiality for wedges of spheres

The polynomiality statement of Theorem 1.5 will follow from the more refined main result of this section. Let Σ\Sigma denote the category of finite sets and bijections999In the representation stability literature this category is called FB., with skeleton the finite sets 𝐧={1,2,,n}\mathbf{n}=\{1,2,\ldots,n\}.

Theorem 5.2.

Let XX be a finite wedge of spheres, possibly of different dimensions, and consider the collision filtration on Hc(F(X,n))H_{c}^{*}(F(X,n)) (defined in 4.5). Its associated graded quotients

grpFHcp+(F(X,n))=FpHcp+(F(X,n))/Fp1Hcp+(F(X,n))\operatorname{gr}_{p}^{F}H_{c}^{p+*}(F(X,n))=F_{p}H_{c}^{p+*}(F(X,n))/F_{p-1}H_{c}^{p+*}(F(X,n))

admit the following algebraic description.

There exists a functor in two variables Ψp:Σ×grVectgrVect\Psi^{p}\colon\Sigma\times\operatorname{grVect}_{\mathbb{Q}}\to\operatorname{grVect}_{\mathbb{Q}} such that Ψp(𝐧,)\Psi^{p}(\mathbf{n},-) is a polynomial functor of degree npn-p with a natural 𝔖n\mathfrak{S}_{n}-action, and such that for any finite wedge of spheres X=iISdiX=\bigvee_{i\in I}{S^{d_{i}}} there is a natural isomorphism of graded 𝔖n\mathfrak{S}_{n}-representations

grpFHcp+(F(X,n))=Ψp(𝐧,H~(X)).\operatorname{gr}_{p}^{F}H_{c}^{p+\ast}(F(X,n))=\Psi^{p}(\mathbf{n},\tilde{H}^{*}(X)).
Note 5.3.

Let us reiterate that while the two inputs of Ψp\Psi^{p} are different types of objects, the equivalence of categories mentioned in Proposition 2.4 relates such functors and bi-functors Σ×ΣgrVect\Sigma\times\Sigma\to\operatorname{grVect}_{\mathbb{Q}} as well as grVect×grVectgrVect\operatorname{grVect}_{\mathbb{Q}}\times\operatorname{grVect}_{\mathbb{Q}}\to\operatorname{grVect}_{\mathbb{Q}} taking inputs of the same type. We prefer the presentation of Ψp\Psi^{p} given here as it lends itself well to a simple geometric interpretation.

A key step in the proof of this theorem is the following lemma. Recall that we defined AWA_{W} to be the square-zero algebra W\mathbb{Q}\oplus W, with trivial multiplication on WW. Furthermore, the CE-complex CCE(AWSLie)(n)C^{CE}_{*}(A_{W}\otimes\operatorname{S}\operatorname{Lie})(n) is bigraded and its filtration by columns is denoted 𝔉\mathfrak{F}_{\bullet}. Thus the graded quotient grp𝔉\operatorname{gr}^{\mathfrak{F}}_{-p} is exactly the pp-th column of the bicomplex.

Lemma 5.4.

Let WW be a graded vector space and fix degree p0p\geq 0. Then for all nn\in\mathbb{N} the functor sending WW to the arity nn term of the Chevalley–Eilenberg homology grp𝔉HCE(AWSLie)(n)\operatorname{gr}_{-p}^{\mathfrak{F}}H^{CE}_{\ast}(A_{W}\otimes\operatorname{S}\operatorname{Lie})(n) is a polynomial functor of degree pp taking values in graded vector spaces.

Proof.

First, the construction WAWSLieW\mapsto A_{W}\otimes\operatorname{S}\operatorname{Lie} produces a twisted dg Lie algebra that in every arity is a polynomial functor of degree 11. Clearly this construction is functorial in WW, i.e. linear maps WWW\to W^{\prime} induce morphisms of twisted dg Lie algebras.

Second, for a Lie algebra 𝔤\mathfrak{g} in symmetric sequences, the pp-th column of the CE-bicomplex is defined as a natural quotient of 𝔤p\mathfrak{g}^{\otimes p}. In every arity nn this expression is given up to degree shifts by

𝔤p=B1Bp=𝐧𝔤[B1]𝔤[Bp],\mathfrak{g}^{\otimes p}=\bigoplus_{B_{1}\coprod\ldots\coprod B_{p}=\mathbf{n}}\mathfrak{g}[B_{1}]\otimes\ldots\otimes\mathfrak{g}[B_{p}],

that is, a tensor product of exactly pp terms from 𝔤\mathfrak{g}. Thus since every one of the 𝔤[k]\mathfrak{g}[k]’s is a polynomial functor of degree 11, then the CE-complex in homological degree pp is polynomial of degree pp.

Since the category of polynomial functors is abelian, it only remains to note that the CE-differentials respect the polynomial functor structure, i.e. that they are natural transformations in WW. This follows from the general fact that the CE-complex CCE(L)C_{-\ast}^{CE}(L) is functorial in its input, the Lie algebra LL.

Proof of Theorem 5.2.

Recall that in Section 4.1 we show that the associated graded of the collision filtration on Hc(F(X,n))H^{*}_{c}(F(X,n)) is computed, after suitable regrading, by the Chevalley–Eilenberg (CE) homology of a twisted Lie algebra, naturally in XX. Explicitly, there is an isomorphism of graded 𝔖n\mathfrak{S}_{n}-representations

grpFHc(F(X,n))grpn𝔉HCE(H(X)SLie(n)).\operatorname{gr}_{p}^{F}H^{*}_{c}(F(X,n))\cong\operatorname{gr}_{p-n}^{\mathfrak{F}}H^{CE}_{\ast}(H^{*}(X)\otimes\operatorname{S}\operatorname{Lie}(n)). (38)

Thus the polynomiality statement reduces to one about this CE-homology.

Furthermore, the cohomology algebra of a wedge of spheres is the square-zero algebra H(X)1H~(X)H^{*}(X)\cong\mathbb{Q}1\oplus\tilde{H}^{*}(X). Therefore the claim of Theorem 5.2 follows from the last lemma. ∎

The polynomiality result thus proved has many consequences for the cohomology Hc(F(X,n))H^{*}_{c}(F(X,n)), but the converse also holds: we next use the cell structure on F(X,n)+F(X,n)^{+} from §4.3 to constrain the nonzero ‘coefficients’ of the polynomial functor and bound its degree.

Proposition 5.5.

For every nn\in\mathbb{N} the polynomial functors Ψp(𝐧,):grVectgrVect\Psi^{p}(\mathbf{n},-)\colon\operatorname{grVect}_{\mathbb{Q}}\to\operatorname{grVect}_{\mathbb{Q}} from the previous theorem have the following properties:

  1. 1.

    Ψp(𝐧,)=0\Psi^{p}(\mathbf{n},-)=0 for all p>n1p>n-1.

  2. 2.

    Ψp(𝐧,W)\Psi^{p}(\mathbf{n},W) decomposes as

    Ψp(𝐧,W)m=np1npΦp[n,m]𝔖mWm\Psi^{p}(\mathbf{n},W)\cong\bigoplus_{m=n-p-1}^{n-p}\Phi^{p}[n,m]\otimes_{\mathfrak{S}_{m}}W^{\otimes m}

    for some 𝔖n×𝔖m\mathfrak{S}_{n}\times\mathfrak{S}_{m}-representations Φp[n,m]\Phi^{p}[n,m]. In particular the polynomial functor Ψp(𝐧,)\Psi^{p}(\mathbf{n},-) has degree npn-p and only two homogeneous terms.

In the ‘leading’ special case p=0p=0, the two nontrivial terms Φ0[n,n]\Phi^{0}[n,n] and Φ0[n,n1]\Phi^{0}[n,n-1] are given as follows.

  1. 3.

    The coefficient Φ0[n,n]\Phi^{0}[n,n] is the "diagonal" representation

    Φ0[n,n]=λnχλχλ.\Phi^{0}[n,n]=\bigoplus_{\lambda\vdash n}{\chi_{\lambda}\boxtimes\chi_{\lambda}}.
  2. 4.

    The coefficient Φ0[n,n1]\Phi^{0}[n,n-1] is

    Φ0[n,n1]=χ(1n)χ(1n1).\Phi^{0}[n,n-1]=\chi_{(1^{n})}\boxtimes\chi_{(1^{n-1})}.

We further note that the highest filtration terms not covered by this proposition, Ψn1[𝐧,]\Psi^{n-1}[\mathbf{n},-] and Ψn2[𝐧,]\Psi^{n-2}[\mathbf{n},-], are computed completely in §5.3, as they consists only of functors of degree 2\leq 2 and are thus multiples of symmetric and alternating powers.

Proof.

The first statement is simply the claim that the associated graded quotient grpHcp+(F(X,n))\operatorname{gr}^{p}H_{c}^{p+*}(F(X,n)) is trivial for p<0p<0 and pnp\geq n, both cases being clear.

For the remaining statements, we use Proposition 2.6 to deduce that it is enough to consider XX a wedge of gg circles with arbitrarily large gg to uniquely determine the coefficients Φp[n,m]\Phi^{p}[n,m]. Then, Proposition 4.12 shows that Hci(F(X,n))=0H^{i}_{c}(F(X,n))=0 unless i{n1,n}i\in\{n-1,n\}, so its associated graded is similarly 0.

On the other hand, set W:=H~(X)W:=\tilde{H}^{*}(X) and consider the graded vector space

grpn𝔉HCE(AWSLie)(n)=qEp,q\operatorname{gr}_{p-n}^{\mathfrak{F}}H^{CE}_{\ast}(A_{W}\otimes\operatorname{S}\operatorname{Lie})(n)=\bigoplus_{q\in\mathbb{Z}}E^{p,q}

with Ep,qE^{p,q} in grading qq. Lemma 4.3 shows that Ep,qgrpFHcp+q(F(X,n))E^{p,q}\cong\operatorname{gr}_{p}^{F}H^{p+q}_{c}(F(X,n)). In light of the previous paragraph it follows that Ep,q=0E^{p,q}=0 unless p+q{n1,n}p+q\in\{n-1,n\}.

But since WW is concentrated in grading q=1q=1, the polynomial description of the CE-homology as Φp[n,m]𝔖mWm\oplus\Phi^{p}[n,m]\otimes_{\mathfrak{S}_{m}}W^{\otimes m} is graded such that its mm-th summand is placed in grading q=mq=m. It follows that

Φp[n,m]𝔖mWm=Ep,m=0\Phi^{p}[n,m]\otimes_{\mathfrak{S}_{m}}W^{\otimes m}=E^{p,m}=0 (39)

unless p+m{n1,n}p+m\in\{n-1,n\}. The second claim now follows since the above calculation is valid for every g1g\geq 1, and since for gmg\geq m the vanishing in (39) implies that Φp[n,m]=0\Phi^{p}[n,m]=0.

The third statement follows from Proposition 4.8: the nn-th row with grading q=nq=n of the E1E_{1}-page has only one nonzero term E10,n=Tn(g)E_{1}^{0,n}=T^{n}(\mathbb{Q}^{g}), the nn-th tensor power, with the 𝔖n\mathfrak{S}_{n}-action by permutation of the tensor factors. Since this term can support no differentials, it survives to E0,nE_{\infty}^{0,n} unchanged. Schur–Weyl duality gives a decomposition of Tn(g)T^{n}(\mathbb{Q}^{g}) as an 𝔖n\mathfrak{S}_{n}-equivariant polynomial functor, agreeing with the claimed expression for Φ0[n,n]\Phi^{0}[n,n].

Our proof of the last statement is long and technical. We defer it to Section 6.2. ∎

Proposition 5.5 now implies Theorem 1.5, since the associated graded grHc(F(X,n))\operatorname{gr}H^{*}_{c}(F(X,n)) is a sum pΨp(𝐧,)\oplus_{p}\Psi^{p}(\mathbf{n},-) up to cohomological shifts, and for p>np>n the functor Ψp(𝐧,)=0\Psi^{p}(\mathbf{n},-)=0 so only finitely many functors contribute for any given nn.

Remark 5.6.

In Theorem 1.5, the coefficients are the graded vector spaces Φ[n,m]\Phi[n,m], while in Proposition 5.5 the coefficients are the (non-graded) vector spaces Φp[n,m]\Phi^{p}[n,m]. These two objects are related as suggested by the notation: Φ[n,m]\Phi[n,m] decomposes as

Φ[n,m]=p=0mΦp[n,m],\Phi[n,m]=\bigoplus_{p=0}^{m}{\Phi^{p}[n,m]},

with Φp[n,m]\Phi^{p}[n,m] being the part of Φ[n,m]\Phi[n,m] in degree pp.

Remark 5.7 (Powell–Vespa’s polynomiality result).

We bring to the reader’s attention the fact that Powell and Vespa proved in [PV18, Theorem 5] a result that is extremely close to our Theorem 5.2. They show that the Hochschild–Pirashvili homology of the classifying spaces B(Fg)B(F_{g}) with square-zero coefficients form a polynomial functor from the category of finitely generated free groups to graded vector spaces – see §2.1.1.

Recalling that B(Fg)gS1B(F_{g})\simeq\bigvee_{g}S^{1} a wedge of 1-spheres, and that Hochschild–Pirashvili homology with square-zero coefficients is dual to compactly supported cohomology of the configuration spaces (Theorem 1.12), their polynomiality result generalizes our Theorem 1.5 for wedges of circles in that we only work at the associated graded level and with the classical notion of polynomial functors. However, their setup does not include wedges of spheres of higher dimensions as we consider here.

For a first geometric consequence of the polynomiality result we give the following.

Corollary 5.8.

Let XX be a finite wedge of spheres, all with the same dimension dd. Considering the collision filtration on Hci(F(X,n))H^{i}_{c}(F(X,n)), the pp-th graded quotient grpFHci(F(X,n))\operatorname{gr}^{F}_{p}H^{i}_{c}(F(X,n)) is nonzero only when i=dn(d1)pi=dn-(d-1)p or i=d(n1)(d1)pi=d(n-1)-(d-1)p.

In particular, when d3d\geq 3 non-zero graded pieces grpFHci(F(X,n))\operatorname{gr}^{F}_{p}H^{i}_{c}(F(X,n)) appear in distinct cohomological degrees ii, so there are no non-trivial extensions between the graded pieces.

Proof.

From Lemma 5.5 the graded quotient grpFHcp+(F(X,n))\operatorname{gr}^{F}_{p}H_{c}^{p+*}(F(X,n)) is polynomial with coefficients Φp[n,m]0\Phi^{p}[n,m]\neq 0 only for p+m{n1,n}p+m\in\{n-1,n\}. Since W:=H~(X)W:=\tilde{H}^{*}(X) is concentrated in grading q=dq=d, the summand

Φp[n,m]𝔖mWm\Phi^{p}[n,m]\otimes_{\mathfrak{S}_{m}}W^{\otimes m}

is concentrated in grading q=dmq=dm. It follows that grpHp+(F(X,n))\operatorname{gr}^{p}H^{p+*}(F(X,n)) is nontrivial only in grading p+dmp+dm where p+m{n1,n}p+m\in\{n-1,n\}. Setting i=p+dmi=p+dm and substituting m{n1p,np}m\in\{n-1-p,\ n-p\} gives the claim.

For the second part of the claim note that the equations dn(d1)p=dn(d1)pdn-(d-1)p=dn-(d-1)p^{\prime} and d(n1)(d1)p=d(n1)(d1)pd(n-1)-(d-1)p=d(n-1)-(d-1)p^{\prime} immediately give p=pp=p^{\prime} when d>1d>1, while the equation dn(d1)p=d(n1)(d1)pdn-(d-1)p=d(n-1)-(d-1)p^{\prime} gives d=(d1)(pp)d=(d-1)(p^{\prime}-p), which has no integral solution for d3d\geq 3 since the last equation means that d1d-1 is a divisor of dd, but in that case d1d-1 is greater than 11 and coprime with dd. ∎

We conclude this section by relating the collision filtration with more familiar filtrations defined on Hochschild–Pirashvili homology. Consider the contravariant functor from finitely generated free groups

FgHc(F(i=1gS1,n))F_{g}\mapsto H^{*}_{c}\left(F(\vee_{i=1}^{g}S^{1},n)\right) (40)

discussed in §4.2. Djament–Vespa [DV15] define a filtration on functors of this sort, whose associated graded quotients factor through the abelianization FggF_{g}\mapsto\mathbb{Z}^{g} and are polynomial in the classical sense. They call this the polynomial filtration, and [PV18, Theorem 17.8] shows that it agrees with another natural filtration – Pirashvili’s so called Hodge filtration on Hochschild–Pirashvili homology. Adapting the filtrations to the contravariant setting of interest here, we have the following.

Corollary 5.9 (Coincidence of filtrations).

For XX a wedge of circles, the collision filtration on Hc(F(X,n))H^{*}_{c}(F(X,n)) coincides with the polynomial filtration of contravariant functors.

Proof.

The collision spectral sequence along with Theorem 5.2 give a natural isomorphism grpFHci(F(X,n))mΦp[n,m]𝔖m[H~(X)m]ip\operatorname{gr}_{p}^{F}H^{i}_{c}(F(X,n))\cong\oplus_{m}\Phi^{p}[n,m]\otimes_{\mathfrak{S}_{m}}\left[\tilde{H}^{*}(X)^{\otimes m}\right]^{i-p}, where [W]q[W]^{q} is the qq-graded part of a graded vector space WW. Indeed Φp[n,m]\Phi^{p}[n,m] is concentrated in grading pp, so H~(X)m\tilde{H}^{*}(X)^{\otimes m} must contribute to grading ipi-p. Since H~(X)\tilde{H}^{*}(X) is concentrated in grading 11, it means that only m=ipm=i-p contributes non-trivially.

In other words, for every ii the functor Hci(F(X,n))H^{i}_{c}(F(X,n)) is filtered by the collision filtration, and the pp-th graded factor is a homogeneous polynomial functor of degree ipi-p. The dual version for contravariant functors of [PV18, Remark 6.10] is stating exactly that such a filtration is unique, and it is the polynomial filtration. ∎

5.2 Schur functor multiplicity

One can make sense of the polynomial functor structure of grHc(F(X,n))\operatorname{gr}H^{*}_{c}(F(X,n)) geometrically using the following fact. Let [X,Y][X,Y] denote the set of homotopy classes of maps between topological spaces XX and YY.

Lemma 5.10.

Let XX and YY be wedges of gg and hh spheres, respectively, all of equal dimension dd. Then the operations induced by [X,Y][X,Y] on homology give a surjection onto the integer matrix space

Hom(Hd(X;),Hd(Y;))Mg×h().\operatorname{Hom}_{\mathbb{Z}}(H_{d}(X;\mathbb{Z}),H_{d}(Y;\mathbb{Z}))\cong M_{g\times h}(\mathbb{Z}).

In particular, [X,X][X,X] surjects onto the matrix ring Mg()M_{g}(\mathbb{Z}).

Proof.

When dealing with wedges of circles, the claim follows since wedges of circles are classifying spaces of free groups and the abelianization map Hom𝒈𝒓𝒑(Fg,Fh)Mg×h()\operatorname{Hom}_{\bm{grp}}(F_{g},F_{h})\to M_{g\times h}(\mathbb{Z}) is surjective. For general dd one only needs to observe that XX and YY are the reduced suspensions of corresponding wedge of circles. The (d1)(d-1)-fold suspension of an appropriate basepoint preserving map between wedges of circles realizes any prescribed homological action. ∎

From Corollary 4.7 the functor XgrHc(F(X,n))X\mapsto\operatorname{gr}H^{*}_{c}(F(X,n)) factors through H(X)H^{*}(X). This uniquely characterizes the polynomial structure on grHc(F(X,n))\operatorname{gr}H^{*}_{c}(F(X,n)), as explained next.

Corollary 5.11 (Uniqueness).

Let Ψ(𝐧,)\Psi^{\prime}(\mathbf{n},-) be any polynomial functor on graded vector spaces, such that its composition with reduced cohomology XH~(X)X\mapsto\tilde{H}^{*}(X) admits a natural isomorphism

Ψ(𝐧,H~(X))grpHcp+(F(X,n))\Psi^{\prime}(\mathbf{n},\tilde{H}^{*}(X))\cong\operatorname{gr}^{p}H^{p+*}_{c}(F(X,n))

as functors out of the full subcategory of finite wedges of spheres. Then Ψ(𝐧,)Ψp(𝐧,)\Psi^{\prime}(\mathbf{n},-)\cong\Psi^{p}(\mathbf{n},-), the functor from Theorem 5.2. In fact, it is already uniquely determined by its restriction to the subcategory of wedges of dd-dimensional spheres for any fixed d1d\geq 1.

Proof.

By naturality in XX, there is a natural isomorphism Ψp(𝐧,H~())Ψ(𝐧,H~())\Psi^{p}(\mathbf{n},\tilde{H}^{*}(-))\cong\Psi^{\prime}(\mathbf{n},\tilde{H}^{*}(-)). In particular, the two functors agree on every homomorphism H(X)H(Y)H^{*}(X)\to H^{*}(Y) induced by a map of spaces. But since the previous lemma shows that these include all integrally defined homomorphisms, and since the integer lattice is Zariski dense in the space of all linear maps, the two polynomial functors must agree on all linear maps. ∎

A structural consequence of the polynomiality of grHc(F(X,n))\operatorname{gr}H^{*}_{c}(F(X,n)) is that it factors into Schur functors, as discussed in §2.1. Explicitly,

grpHcp+(F(X,n))λΦp[n,λ]𝕊λ(H~(X))\operatorname{gr}^{p}H^{p+*}_{c}(F(X,n))\cong\bigoplus_{\lambda}\Phi^{p}[n,\lambda]\boxtimes\mathbb{S}_{\lambda}\left(\tilde{H}^{*}(X)\right) (41)

where Φp[n,λ]\Phi^{p}[n,\lambda] is some 𝔖n\mathfrak{S}_{n}-representation and 𝕊λ()\mathbb{S}_{\lambda}(-) is a Schur functor. Note that Φ[n,λ]\Phi[n,\lambda] is the χλ\chi_{\lambda}-multiplicity space in the 𝔖m\mathfrak{S}_{m}-representation Φ[n,m]\Phi[n,m] from the introduction.

Definition 5.12.

The 𝔖n\mathfrak{S}_{n}-representation Φp[n,λ]\Phi^{p}[n,\lambda] appearing in (41) is the equivariant multiplicity of the Schur functor 𝕊λ\mathbb{S}_{\lambda} in grpHcp+(F(X,n))\operatorname{gr}^{p}H^{p+*}_{c}(F(X,n)).

Understanding the cohomology grHc(F(X,n))\operatorname{gr}H^{*}_{c}(F(X,n)) amounts to describing these equivariant multiplicities, e.g. giving their characters for all pp.

Observation 5.13 (Genus bound principle).

Let λm\lambda\vdash m be a partition with \ell parts. Then given X=i=1gSdX=\bigvee_{i=1}^{g}S^{d}, the 𝔖n\mathfrak{S}_{n}-equivariant multiplicity of the Schur functor 𝕊λ(H~(X))\mathbb{S}_{\lambda}(\tilde{H}(X)) under the natural action by the monoid of homotopy classes [X,X][X,X] on Hc(F(X,n))H_{c}^{*}(F(X,n)) is independent of gg once gg\geq\ell. In particular, this multiplicity could be read off from XX a wedge of exactly \ell spheres.

Proof.

Let λ\lambda be a partition with \ell parts, that is λ=(λ1,λ2,,λ)\lambda=(\lambda_{1},\lambda_{2},\ldots,\lambda_{\ell}), and consider XX a wedge of gg\geq\ell spheres of dimension dd.

Consider the surjection [X,X]End(g)[X,X]\twoheadrightarrow\operatorname{End}(\mathbb{Z}^{g}) given by the homological action. The Schur functors 𝕊μ(g)\mathbb{S}_{\mu}(\mathbb{Q}^{g}) for partitions μ\mu with g\leq g parts are distinct nonzero irreducible representations of End(g)\operatorname{End}(\mathbb{Z}^{g}), differentiated e.g. by their characters on diagonal matrices: these are evaluations of the respective Schur polynomial sμ(x1,,xg)s_{\mu}(x_{1},\ldots,x_{g}) at the diagonal entries. In particular, 𝕊λ(H~(X))\mathbb{S}_{\lambda}(\tilde{H}^{*}(X)) is a nonzero irreducible representation of [X,X][X,X], distinct from all other Schur functors appearing in Decomposition (41). Thus by Schur’s lemma, there is an 𝔖n\mathfrak{S}_{n}-equivariant isomorphism

Φp[n,λ]Hom[X,X](𝕊λ(H~(X)),grpHcp+(F(X,n)))\Phi^{p}[n,\lambda]\cong\operatorname{Hom}_{[X,X]}\left(\mathbb{S}_{\lambda}(\tilde{H}^{*}(X)),\operatorname{gr}^{p}H^{p+*}_{c}(F(X,n))\right) (42)

determining the equivariant multiplicity independently of gg. ∎

Swapping even spheres for odd spheres has the effect of conjugating the partition up to grading shifts

𝕊λ(H~(X)){𝕊λ(g) if X=i=1gS2d𝕊λ(g) if X=i=1gS2d+1.\mathbb{S}_{\lambda}(\tilde{H}^{*}(X))\cong\begin{cases}\mathbb{S}_{\lambda}(\mathbb{Q}^{g})&\text{ if }X=\bigvee_{i=1}^{g}S^{2d}\\ \mathbb{S}_{{\lambda}^{\ast}}(\mathbb{Q}^{g})&\text{ if }X=\bigvee_{i=1}^{g}S^{2d+1}.\end{cases} (43)

Thus the notion of ‘simple’ partitions, detectable by wedges of few spheres, includes partitions λ\lambda such that either λ\lambda or its conjugate λ{\lambda}^{\ast} have few parts.

Corollary 5.14.

Let XX be a wedge of gg spheres of dimension dd and fix a partition λ\lambda. Then under the Mg()M_{g}(\mathbb{Z})-action induced by [X,X][X,X] on homology, the 𝔖n\mathfrak{S}_{n}-equivariant multiplicity HomMg()(𝕊λ(g),grpHcp+(F(X,n)))\operatorname{Hom}_{M_{g}(\mathbb{Z})}\left(\mathbb{S}_{\lambda}(\mathbb{Q}^{g}),\operatorname{gr}^{p}H^{p+*}_{c}(F(X,n))\right) is isomorphic to

{Φp[n,λ]d even and λ has g partsΦp[n,λ]d odd and λ has g parts0otherwise.\begin{cases}\Phi^{p}[n,\lambda]&d\text{ even and $\lambda$ has $\leq g$ parts}\\ \Phi^{p}[n,{\lambda}^{\ast}]&d\text{ odd and ${\lambda}^{\ast}$ has $\leq g$ parts}\\ 0&\text{otherwise}.\end{cases} (44)
Remark 5.15.

Proposition 5.5 above showed that Φp[n,λ]0\Phi^{p}[n,\lambda]\neq 0 only if p+|λ|{n1,n}p+|\lambda|\in\{n-1,n\}. In other words, the only equivariant multiplicities to compute are the ones with p=n1|λ|p=n-1-|\lambda| and n|λ|n-|\lambda|.

The first examples of ‘simple’ partitions are λ=(m)\lambda=(m) and (1m)(1^{m}). We utilize the principle thus outlined in the next section.

5.3 Symmetric and alternating powers

We now compute the 𝔖n\mathfrak{S}_{n}-equivariant multiplicity of Symm()\operatorname{Sym}^{m}(-) and Λm()\Lambda^{m}(-) occurring in grpHcp+(F(,n))\operatorname{gr}^{p}H_{c}^{p+*}(F(-,n)). The determination are given in terms of other geometric objects whose homology is well-understood, and they will prove Theorem 1.8.

Since symmetric and alternating powers are 𝕊λ\mathbb{S}_{\lambda} for partition λ\lambda with only one row or column, their equivariant multiplicity in grHc(F(X,n))\operatorname{gr}H^{*}_{c}(F(X,n)) are determined by configurations on a single sphere. These spaces have been studied extensively, most notably [FZ00] computed the integral cohomology rings of F(Sd,n)F(S^{d},n) for all (d,n)(d,n).

Our calculations for Symm\operatorname{Sym}^{m} and Λm\Lambda^{m} follow a very similar pattern: projecting configurations on a sphere to a moduli space of such configurations. Let us begin with the simpler case of alternating powers.

Proposition 5.16 (Alternating powers).

For the partition λ=(1m){\lambda}^{\ast}=(1^{m}) the Schur functor 𝕊λ=Λm\mathbb{S}_{{\lambda}^{\ast}}=\Lambda^{m}. The equivariant multiplicity Φp[n,(1m)]\Phi^{p}[n,(1^{m})] of Λm\Lambda^{m} in the functor Ψp(𝐧,)\Psi^{p}(\mathbf{n},-) is

{H2(nm)(F(3,n1))sgnnif p=nmH2(nm1)(F(3,n1))sgnnif p=nm10otherwise,\cong\begin{cases}H_{2(n-m)}(F(\mathbb{R}^{3},n-1))\otimes\operatorname{sgn}_{n}&\text{if }p=n-m\\ H_{2(n-m-1)}(F(\mathbb{R}^{3},n-1))\otimes\operatorname{sgn}_{n}&\text{if }p=n-m-1\\ 0&\text{otherwise},\end{cases} (45)

where 𝔖n\mathfrak{S}_{n} acts on F(3,n1)F(\mathbb{R}^{3},n-1) via the identification with F(SU(2),n)/SU(2)F(SU(2),n)/SU(2).

These 𝔖n\mathfrak{S}_{n}-representations are the Whitehouse modules up to sign, see [ER19].

Proof.

Let λ=(m)\lambda=(m) so that λ=(1m){\lambda}^{\ast}=(1^{m}) and consider configurations on S3S^{3}. Since λ\lambda has only one part, Corollary 5.14 identifies Φp[n,λ]\Phi^{p}[n,{\lambda}^{\ast}] as the equivariant multiplicity of Symm()\operatorname{Sym}^{m}(\mathbb{Q}) in grpHcp+(F(S3,n))\operatorname{gr}^{p}H^{p+*}_{c}(F(S^{3},n)) as a representation of π0(Maps(S3,S3))\pi_{0}(\operatorname{Maps}({S^{3}},{S^{3}}))\cong\mathbb{Z}. Here aa\in\mathbb{Z} acts on Symm()\operatorname{Sym}^{m}(\mathbb{Q}) as multiplication by ama^{m}. Thus we proceed by computing Hc(F(S3,n))H^{*}_{c}(F(S^{3},n)).

First, since F(S3,n)F(S^{3},n) is a manifold, Poincaré duality gives an 𝔖n\mathfrak{S}_{n}-equivariant isomorphism

Hc3ni(F(S3,n))sgnnHi(F(S3,n))H^{3n-i}_{c}(F(S^{3},n))\otimes\operatorname{sgn}_{n}\cong H_{i}(F(S^{3},n)) (46)

where the additional sign comes from the induced 𝔖n\mathfrak{S}_{n}-action on the orientation bundle of (S3)n(S^{3})^{n}.

The cohomology of F(S3,n)F(S^{3},n) was computed in [FZ00], but we seek a different description. Thinking of S3S^{3} as the group SU(2)SU(2), the quotient by the diagonal action

F(SU(2),n)F(SU(2),n)/SU(2)F(SU(2),n)\to F(SU(2),n)/SU(2) (47)

is a trivial SU(2)SU(2)-principal bundle. This is furthermore an 𝔖n\mathfrak{S}_{n}-equivariant map. Since SU(2)SU(2) has homology only in degrees 0 and 33, there is an 𝔖n\mathfrak{S}_{n}-equivariant isomorphism101010For 𝔖n\mathfrak{S}_{n}-equivariance one also needs to observe that the 𝔖n\mathfrak{S}_{n}-action on the SU(2)SU(2) fibers is homologically trivial. Indeed, one can check that the 𝔖n\mathfrak{S}_{n}-action commutes with the left SU(2)SU(2) action. Thus it must act on every copy of SU(2)SU(2) as right multiplication by a fixed matrix. But since SU(2)SU(2) is connected, such multiplication is homotopically trivial.

Hi(F(S3,n))Hi(F(SU(2),n)/SU(2))Hi3(F(SU(2),n)/SU(2)).\displaystyle\begin{split}H_{i}(F(S^{3},n))\quad\cong\quad&H_{i}(F(SU(2),n)/SU(2))\\ &\oplus H_{i-3}(F(SU(2),n)/SU(2)).\end{split} (48)

Via the identification SU(2){1}S3{N}3SU(2)\setminus\{1\}\cong S^{3}\setminus\{N\}\cong\mathbb{R}^{3} one gets a homeomorphism F(SU(2),n)/SU(2)F(3,n1)F(SU(2),n)/SU(2)\cong F(\mathbb{R}^{3},n-1) by mapping

(x1,,xn)(x11x2,,x11xn).(x_{1},\ldots,x_{n})\mapsto(x_{1}^{-1}x_{2},\ldots,x_{1}^{-1}x_{n}).

It is furthermore well-known that, for n2n\geq 2, H(F(3,n1))H_{*}(F(\mathbb{R}^{3},n-1)) is concentrated in grading 2k2k for 0kn20\leq k\leq n-2 (see [Coh76, Lemma 6.2]). Therefore the homology H(F(S3,n)))H_{*}(F(S^{3},n))) coincides with H(F(3,n1))H_{*}(F(\mathbb{R}^{3},n-1)) in even degrees, and with H3(F(3,n1))H_{*-3}(F(\mathbb{R}^{3},n-1)) in odd ones. Under Poincaré duality these become Hc3nH^{3n-*}_{c} in even and odd codimension respectively.

Let us match (48) with the collision filtration. By Proposition 5.5 the only nontrivial terms in grpHcp+(F(S3,n))\operatorname{gr}^{p}H^{p+*}_{c}(F(S^{3},n)) are those in grading =3q*=3q where p+q{n1,n}p+q\in\{n-1,n\}. If p+q=np+q=n then the term in degree p+3qp+3q has even codimension 3n(p+3q)=2(nq)3n-(p+3q)=2(n-q), and similarly p+q=n1p+q=n-1 implies odd codimension 2(nq)+12(n-q)+1. Thus by parity of dimensions the collision filtration has no extensions:

Hc3ni(F(S3,n))={grkHc3n2kif i=2k evengrk1Hc3n2k1if i=2k+1 oddH^{3n-i}_{c}(F(S^{3},n))=\begin{cases}\operatorname{gr}^{k}H^{3n-2k}_{c}&\text{if $i=2k$ even}\\ \operatorname{gr}^{k-1}H^{3n-2k-1}_{c}&\text{if $i=2k+1$ odd}\end{cases} (49)

Now recall that grpHcp+(F(S3,n))mΦp[n,(1m)]Symm()\operatorname{gr}^{p}H^{p+*}_{c}(F(S^{3},n))\cong\oplus_{m}\Phi^{p}[n,(1^{m})]\boxtimes\operatorname{Sym}^{m}(\mathbb{Q}) where the mm-th symmetric power has grading =3m*=3m. Considering the case of even codimension first, with k=pk=p and 3m==3n3p3m=*=3n-3p we have an equivariant isomorphism

grpHc3n2p=Hc3n2pH2p(F(3,n1))sgnn\operatorname{gr}^{p}H_{c}^{3n-2p}=H_{c}^{3n-2p}\cong H_{2p}(F(\mathbb{R}^{3},n-1))\otimes\operatorname{sgn}_{n}

which equivariantly identifies the degree m=npm=n-p summand

Φp[n,(1np)]Symnp()H2p(F(3,n1))sgnn.\Phi^{p}[n,(1^{n-p})]\boxtimes\operatorname{Sym}^{n-p}(\mathbb{Q})\cong H_{2p}(F(\mathbb{R}^{3},n-1))\otimes\operatorname{sgn}_{n}. (50)

Considered as 𝔖n\mathfrak{S}_{n}-representations, the term Symnp()\operatorname{Sym}^{n-p}(\mathbb{Q}) is a trivial 11-dimensional representation, and thus the first case of the proposition follows.

The odd codimension case is similar: taking p=kp=k and grading 3m==3n3p3m=*=3n-3p we have

grp1Hc3n2p1=Hc3n2p1H2(p1)(F(3,n1))sgnn.\operatorname{gr}^{p-1}H^{3n-2p-1}_{c}=H^{3n-2p-1}_{c}\cong H_{2(p-1)}(F(\mathbb{R}^{3},n-1))\otimes\operatorname{sgn}_{n}.

This gives the equivariant identification in degree m=npm=n-p

Φp1[n,(1np)]Symnp()H2(p1)(F(3,n1))sgnn\Phi^{p-1}[n,(1^{n-p})]\boxtimes\operatorname{Sym}^{n-p}(\mathbb{Q})\cong H_{2(p-1)}(F(\mathbb{R}^{3},n-1))\otimes\operatorname{sgn}_{n} (51)

which produces the second case of the theorem. The remaining cases vanish due to Proposition 5.5. ∎

Let us now consider symmetric powers. In this case the collision filtration exhibits an extension, and we identify its terms using Deligne’s theory of weights.

Proposition 5.17 (Symmetric powers).

Let n3n\geq 3. For the partition λ=(m)\lambda=(m) the Schur functor 𝕊λ=Symm\mathbb{S}_{\lambda}=\operatorname{Sym}^{m}. The equivariant multiplicity Φp[n,(m)]\Phi^{p}[n,(m)] of Symm\operatorname{Sym}^{m} in the functor Ψp(𝐧,)\Psi^{p}(\mathbf{n},-) is

{Hnm(0,n)if p=nmHnm2(0,n)if p=nm10otherwise,\cong\begin{cases}H_{n-m}(\mathcal{M}_{0,n})&\text{if }p=n-m\\ H_{n-m-2}(\mathcal{M}_{0,n})&\text{if }p=n-m-1\\ 0&\text{otherwise},\end{cases} (52)

where 0,n\mathcal{M}_{0,n} is the moduli space of genus 0 algebraic curves with nn marked points.

Proof.

Proceeding as in the previous case, Corollary 5.14 identifies Ψp[n,(m)]\Psi^{p}[n,(m)] with the multiplicity of Symm()\operatorname{Sym}^{m}(\mathbb{Q}) in grpHcp+(F(S2,n))\operatorname{gr}^{p}H^{p+*}_{c}(F(S^{2},n)) as a representation of the group π0(Maps(S2,S2))\pi_{0}(\operatorname{Maps}({S^{2}},{S^{2}}))\cong\mathbb{Z}. We therefore seek to understand Hc(F(S2,n))H^{*}_{c}(F(S^{2},n)).

Since F(S2,n)F(S^{2},n) is a manifold, Poincaré duality gives an 𝔖n\mathfrak{S}_{n}-equivariant isomorphism

Hc2n(F(S2,n))H(F(S2,n)).H^{2n-*}_{c}(F(S^{2},n))\cong H_{*}(F(S^{2},n)).

Thinking of S2S^{2} as P1\mathbb{C}P^{1}, the group PSL2()PSL_{2}(\mathbb{C}) acts 3-transitively and [FZ00, Theorem 2.1] shows that the quotient map

F(P1,n)0,nF(\mathbb{C}P^{1},n)\to\mathcal{M}_{0,n}

is a trivial PSL2()PSL_{2}(\mathbb{C})-principal bundle. This projection is clearly 𝔖n\mathfrak{S}_{n}-equivariant. Moreover, all spaces and maps involved are algebraic, so their homology is equipped with a natural Hodge structure.

Since the rational cohomology of PSL2()PSL_{2}(\mathbb{C}) is the same as that of S32{0}S^{3}\simeq\mathbb{C}^{2}\setminus\{0\}, by the same argument as in the case of alternating powers there is an 𝔖n\mathfrak{S}_{n}-equivariant isomorphism

Hi(F(S2,n))Hi(0,n)Hi3(0,n)(2).H_{i}(F(S^{2},n))\cong H_{i}(\mathcal{M}_{0,n})\oplus H_{i-3}(\mathcal{M}_{0,n})\otimes\mathbb{Q}(-2). (53)

It is also known that Hi(0,n)H_{i}(\mathcal{M}_{0,n}) is pure of weight 2i2i as the complement of a hyperplane arrangement. Overall it follows that Hi(F(S2,n))H_{i}(F(S^{2},n)) is mixed of weights 2i2i and 2i22i-2.

We want to compare (53) with the collision filtration. By Proposition 5.5 there is an extension

0griFHc2ni(F(S2,n))Hc2ni(F(S2,n))gri2FHc2ni(F(S2,n))00\to\operatorname{gr}_{i}^{F}H^{2n-i}_{c}(F(S^{2},n))\to H^{2n-i}_{c}(F(S^{2},n))\to\operatorname{gr}_{i-2}^{F}H^{2n-i}_{c}(F(S^{2},n))\to 0 (54)

and we wish to match it with the one in (53). It suffices to show that griFHc2ni(F(S2,n))\operatorname{gr}_{i}^{F}H^{2n-i}_{c}(F(S^{2},n)) is pure of weight 2n2i2n-2i and gri2FHc2ni(F(S2,n))\operatorname{gr}_{i-2}^{F}H^{2n-i}_{c}(F(S^{2},n)) is pure of weight 2n(2i2)2n-(2i-2), for then under Poincaré duality they must match Hi(0,n)H_{i}(\mathcal{M}_{0,n}) and Hi3(0,n)(2)H_{i-3}(\mathcal{M}_{0,n})\otimes\mathbb{Q}(-2) respectively.

Petersen showed in [Pet17, §3.2] that the collision spectral sequence is compatible with Hodge structures, thus it suffices to show the claimed purity on the E1E_{1} page. We will prove generally that E1p,2qE_{1}^{p,2q} has pure weight 2q2q. Indeed, Proposition 4.8 shows that E1p,2qE_{1}^{p,2q} is a sum of H2(P1)qH^{2}(\mathbb{C}P^{1})^{\otimes q}, and thus has the claimed pure weight.

Now grpFHcp+(F(S2,n))mΦp[n,(m)]Symm()\operatorname{gr}_{p}^{F}H^{p+*}_{c}(F(S^{2},n))\cong\oplus_{m}\Phi^{p}[n,(m)]\boxtimes\operatorname{Sym}^{m}(\mathbb{Q}) where the mm-th symmetric power has grading 2m2m. Together with griFHc2ni(F(S2,n))Hi(0,n)\operatorname{gr}_{i}^{F}H^{2n-i}_{c}(F(S^{2},n))\cong H_{i}(\mathcal{M}_{0,n}) it follows that for i=pi=p and 2m==2n2p2m=*=2n-2p we have an isomorphism

Φp[n,(np)]Symnp()Hp(0,n).\Phi^{p}[n,(n-p)]\boxtimes\operatorname{Sym}^{n-p}(\mathbb{Q})\cong H_{p}(\mathcal{M}_{0,n}). (55)

The first case of the theorem follows.

The same argument applied to gri2FHc2ni(F(S2,n))Hi3(0,n)(2)\operatorname{gr}_{i-2}^{F}H^{2n-i}_{c}(F(S^{2},n))\cong H_{i-3}(\mathcal{M}_{0,n})\otimes\mathbb{Q}(-2) gives the second case. And the vanishing in all other cases follows from Proposition 5.5. ∎

Remark 5.18 (End(1)\operatorname{End}(\mathbb{P}^{1})-action on 0,n\mathcal{M}_{0,n}).

A consequence of the calculation in the proof is a description of the π0(End(1))\pi_{0}(\operatorname{End}(\mathbb{P}^{1}))-action on H(0,n)H_{*}(\mathcal{M}_{0,n}), defined via Poincaré duality H(0,n)Hc(0,n)H_{*}(\mathcal{M}_{0,n})\cong H^{*}_{c}(\mathcal{M}_{0,n}) with End(1)\operatorname{End}(\mathbb{P}^{1}) acting on the cohomology. Explicitly, (55) shows that a degree kk map on 1\mathbb{P}^{1} acts on Hp(0,n)H_{p}(\mathcal{M}_{0,n}) as multiplication by knpk^{n-p}.

5.3.1 Explicit characters

The 𝔖n\mathfrak{S}_{n}-representations arising in the previous section can be understood combinatorially: let us recall their characters. For every 𝔖n\mathfrak{S}_{n}-representation WW, let chn(W)Λ\operatorname{ch}_{n}(W)\in\Lambda denote the Frobenius characteristic of WW,

chn(W):=λnχW(λ)zλpλ\operatorname{ch}_{n}(W):=\sum_{\lambda\vdash n}\frac{\chi_{W}(\lambda)}{z_{\lambda}}p_{\lambda}

where pλp_{\lambda} are the power-sum symmetric functions, χW(λ)\chi_{W}(\lambda) is the character value of WW on a permutation of cycle type λ\lambda, and zλ=imimi!z_{\lambda}=\prod i^{m_{i}}m_{i}! for λ=(1m1,2m2,,nmn)\lambda=(1^{m_{1}},2^{m_{2}},\ldots,n^{m_{n}}).

The equivariant Poincaré polynomial for H(0,n)H_{\ast}(\mathcal{M}_{0,n}) was computed by Getzler in [Get95, Theorem 5.7], where he gave the following formula in the ring of symmetric functions. Below, raising one symmetric function to the power of another fgf^{g} is interpreted using the plethystic exponential and logarithm of Getzler–Kapranov [GK98]

fg:=Exp(gLog(f)).f^{g}:=\operatorname{Exp}(g\cdot\operatorname{Log}(f)).
Proposition 5.19.

The characters of the 𝔖n\mathfrak{S}_{n}-representations Hi(0,n)H_{i}(\mathcal{M}_{0,n}) are encoded by the generating function Cht(𝐦)=n=3i=0(t)ichn(Hi(0,n))\operatorname{Ch}_{t}(\mathbf{m})=\sum_{n=3}^{\infty}\sum_{i=0}^{\infty}{(-t)^{i}ch_{n}(H_{i}(\mathcal{M}_{0,n}))} given by

Cht(𝐦)=κ1+tp11t2n=1(1+tnpn)Rn(t)\operatorname{Ch}_{t}(\mathbf{m})=\kappa\frac{1+tp_{1}}{1-t^{2}}\prod_{n=1}^{\infty}{(1+t^{n}p_{n})^{R_{n}(t)}} (56)

where pi=(x1i+x2i+)p_{i}=(x_{1}^{i}+x_{2}^{i}+\ldots) denote the power-sum symmetric functions, κ\kappa is a truncation operator sending the monomials 11, p1p_{1}, p12p_{1}^{2} and p2p_{2} to zero while fixing the other monomials, and

Rn(t)=1ndnμ(n/d)tdR_{n}(t)=\frac{1}{n}\sum_{d\mid n}{\frac{\mu(n/d)}{t^{d}}}

with μ\mu the Möbius function.

The character of Hi(0,n)H_{i}(\mathcal{M}_{0,n}) is thus obtained from the degree nn part of the symmetric function appearing as coefficient for tit^{i} in (56).

A formula for the character of the 𝔖n\mathfrak{S}_{n}-representation Hi(F(3,n1))H_{i}(F(\mathbb{R}^{3},n-1)) was given by Pagaria very recently.

Theorem 5.20 ([Pag22, Corollary 4.11]).

The characters of the 𝔖n\mathfrak{S}_{n}-representations Hi(F(3,n1))H_{i}(F(\mathbb{R}^{3},n-1)) are encoded by Cht(𝐰)=n=1i=0tichn(H2i(F(3,n1)))\operatorname{Ch}_{t}(\mathbf{w})=\sum_{n=1}^{\infty}\sum_{i=0}^{\infty}{t^{i}ch_{n}(H_{2i}(F(\mathbb{R}^{3},n-1)))} where

Cht(𝐰)=11t((1tp1)t1t1).\operatorname{Ch}_{t}(\mathbf{w})=\frac{1}{1-t}\left((1-t\cdot p_{1})^{\frac{t-1}{t}}-1\right). (57)

5.4 A lower bound

The Euler characteristic of the collision spectral sequence (30) gives a concrete lower bound on grHc(F(X,n))\operatorname{gr}H^{*}_{c}(F(X,n)). Let us focus on configurations on wedges of circles: X=i=1gS1X=\bigvee_{i=1}^{g}S^{1}. In this section we will use the shorthand HcH_{c}^{*} to denote Hc(F(X,n))H_{c}^{*}(F(X,n)).

Lemma 5.21 (Lower bound).

Decompose the terms of the Chevalley–Eilenberg complex of the twisted Lie algebra (V[1])SLie(\mathbb{Q}\oplus V[-1])\otimes\operatorname{S}\operatorname{Lie} as

Symp((V[1])SLie[1])(n)q=1nMp,q𝔖qVq\operatorname{Sym}^{p}\left((\mathbb{Q}\oplus V[-1])\otimes\operatorname{S}\operatorname{Lie}[1]\right)(n)\cong\bigoplus_{q=1}^{n}M^{p,q}\otimes_{\mathfrak{S}_{q}}V^{\otimes q} (58)

for 𝔖nMp,q𝔖q\mathfrak{S}_{n}\curvearrowright M^{p,q}\curvearrowleft\mathfrak{S}_{q} explicitly computable (though complicated) 𝔖n×𝔖q\mathfrak{S}_{n}\times\mathfrak{S}_{q}-bimodules.

Fix q1q\geq 1 and suppose the irreducible χλχμ\chi_{\lambda}\boxtimes\chi_{\mu} appears in the virtual representation

p=0nq(1)p+qMp,q\sum_{p=0}^{n-q}(-1)^{p+q}M^{p,q}

with (signed) multiplicity mm. When mm has sign (1)n(-1)^{n}, the top cohomology grHcn\operatorname{gr}H^{n}_{c} contain at least |m||m| copies of χλ𝕊μ(H~1(X))\chi_{\lambda}\otimes\mathbb{S}_{\mu}(\tilde{H}^{1}(X)). Otherwise, grHcn1\operatorname{gr}H^{n-1}_{c} contains at least |m||m| copies of χλ𝕊μ(H~1(X))\chi_{\lambda}\otimes\mathbb{S}_{\mu}(\tilde{H}^{1}(X)).

Explicit computations are easily carried out on a home computer. Some consequences are given below.

Proof.

The E1E_{1}-page of the collision spectral sequence (30) is given by the Chevalley–Eilenberg complex on H(X)SLieH^{*}(X)\otimes\operatorname{S}\operatorname{Lie}, and H(X)=H1(X)[1]H^{*}(X)=\mathbb{Q}\oplus H^{1}(X)[-1]. Since the Euler characteristic can be computed at every page of a spectral sequence, the degree qq polynomial subfunctor of (grHcn)(grHcn1)(\operatorname{gr}H^{n}_{c})-(\operatorname{gr}H^{n-1}_{c}) has the form

(p=0nq(1)p+qMp,q)𝔖qH~1(X)q.\Big{(}\sum_{p=0}^{n-q}(-1)^{p+q}M^{p,q}\Big{)}\otimes_{\mathfrak{S}_{q}}\tilde{H}^{1}(X)^{\otimes q}. (59)

As this Euler characteristic is in fact the difference between two genuine representations, the claim follows. ∎

Remark 5.22.

Together with the exact multiplicities of symmetric and alternating powers from §5.3, the lower bound produced in this way has thus far proved to be a good approximation for the true cohomology. For example, up to n=11n=11 particles, the ranks our estimates produce capture 90%\sim 90\% of the cohomology.

Forgetting the 𝔖n\mathfrak{S}_{n}-action, the lower bound gives the following explicit estimate on the (nonequivariant) multiplicity of every Schur functor.

Proposition 5.23.

Let λq\lambda\vdash q be any partition, then the Schur functor 𝕊λ\mathbb{S}_{\lambda} appears in (grHcn)(grHcn1)(\operatorname{gr}H^{n}_{c})-(\operatorname{gr}H^{n-1}_{c}) with nonequivariant multiplicity

(1)n1(|s(n1,q)||s(n1,q1)|)dim(χλ)(-1)^{n-1}\big{(}|s(n-1,q)|-|s(n-1,q-1)|\big{)}\dim(\chi_{\lambda})

where s(n,q)s(n,q) is a Stirling number of the first kind and dim(χλ)\dim(\chi_{\lambda}) has a combinatorial description as the number of standard Young tableaux of shape λ\lambda, equivalently given by the hook-length formula.

Proof.

In Proposition 4.8 we noted that E1p,qE_{1}^{p,q} consists of |s(n,np)|(npq)|s(n,n-p)|\cdot\binom{n-p}{q} copies of the tensor power Tq(H~1(X))T^{q}(\tilde{H}^{1}(X)). By Schur–Weyl duality, the tensor power decomposes as

λqχλ𝕊λ(H~1(X)).\bigoplus_{\lambda\vdash q}\chi_{\lambda}\boxtimes\mathbb{S}_{\lambda}(\tilde{H}^{1}(X)).

Thus the Schur functor 𝕊λ\mathbb{S}_{\lambda} appears in the Euler characteristic with multiplicity

p=0nq(1)p+q|s(n,np)|(npq)dim(χλ).\sum_{p=0}^{n-q}(-1)^{p+q}|s(n,n-p)|\cdot\binom{n-p}{q}\cdot\dim(\chi_{\lambda}).

Recalling that the Stirling numbers are defined by the relation (x)n:=x(x1)(xn+1)=s(n,p)xp(x)_{n}:=x(x-1)\ldots(x-n+1)=\sum s(n,p)x^{p}, the above sums simplify by considering the generating function

p,q(1)p|s(n,np)|(npq)xq=(x+1)n.\sum_{p,q}(-1)^{p}|s(n,n-p)|\cdot\binom{n-p}{q}x^{q}=(x+1)_{n}.

Comparing the qq-th coefficients of (x+1)n=(x+1)(x)n1(x+1)_{n}=(x+1)\cdot(x)_{n-1} one arrives at the claimed multiplicity. ∎

Example 5.24.

The partition (2,1)3(2,1)\vdash 3 is the smallest one not accounted for in §5.3. The number of copies of 𝕊(2,1)\mathbb{S}_{(2,1)} counted by the Euler characteristic is

=(1)n12(|s(n1,3)||s(n1,2)|)=(1)n1(n2)![(logne1γ)2O(1)]\begin{split}&=(-1)^{n-1}2\left(|s(n-1,3)|-|s(n-1,2)|\right)\\ &=(-1)^{n-1}(n-2)!\left[{\left(\log\frac{n}{e^{1-\gamma}}\right)}^{2}-O(1)\right]\end{split} (60)

as follows from the well-know comparison between Stirling numbers and harmonic sums (see e.g. [Ada97]). Here γ\gamma is the Euler–Mascheroni constant so that e1γ1.5e^{1-\gamma}\approx 1.5. In particular the multiplicity of 𝕊(2,1)\mathbb{S}_{(2,1)} in grHcn1\operatorname{gr}H^{n-1}_{c} grows super exponentially in nn.

More generally, the estimate |s(n1,q)|(n2)!(q1)!(log(n)+γ)q|s(n-1,q)|\sim\frac{(n-2)!}{(q-1)!}(\log(n)+\gamma)^{q} from [Ada97] shows that for every λ\lambda the multiplicity of 𝕊λ\mathbb{S}_{\lambda} in (grHcn1)(grHcn)(\operatorname{gr}H^{n-1}_{c})-(\operatorname{gr}H^{n}_{c}) is at least on the order of C(n2)!log(n/c)qC(n-2)!\log(n/c)^{q} for some constants CC and cc.

Remark 5.25.

The last estimate is surprising. On the one hand it implies that Schur functors of low degree |λ||\lambda| are hugely more prevalent in the bottom cohomology grHcn1\operatorname{gr}H^{n-1}_{c} than in the top grHcn\operatorname{gr}H^{n}_{c}. On the other hand the presentation in (34) shows that dimHcndimHcn1(n!)ng\dim H^{n}_{c}-\dim H^{n-1}_{c}\sim(n!)^{n^{g}} so the top cohomology HcnH^{n}_{c} is much larger overall. One concludes that the top cohomology contains many more Schur functors of high degree.

To be more explicit, an estimate of Erdös [Erd53] gives that the numbers |s(n,q)||s(n,q)| are monotonically increasing in qq until qlog(n)q\approx\log(n) and then they decrease. This means that 𝕊λ\mathbb{S}_{\lambda} with |λ|log(n)|\lambda|\lesssim\log(n) are more common in grHcn1\operatorname{gr}H^{n-1}_{c}, but all remaining ones are more prevalent in grHcn\operatorname{gr}H^{n}_{c}. It further follows that every Schur functor with |λ|n|\lambda|\leq n does appear in grHc\operatorname{gr}H^{*}_{c}, with only one exception for λ=(2)\lambda=(2), since Sym2\operatorname{Sym}^{2} does not appear in Hc(F(X,3))H^{*}_{c}(F(X,3)).

6 Applications

6.1 Weight 0 cohomology of 2,n\mathcal{M}_{2,n}

Our original motivation for studying the cohomology Hc(F(X,n))H^{*}_{c}(F(X,n)) is its connection with the cohomology of moduli spaces of algebraic curves g,n\mathcal{M}_{g,n}. This relationship comes from the description of the weight 0 part of Hc(g,n)H^{*}_{c}(\mathcal{M}_{g,n}) in terms of tropical geometry, and is manifested most explicitly in genus g=2g=2 as given by the following theorem, which can be found under an equivalent formulation in [BCGY21].

Theorem 6.1 ([BCGY21, Theorem 1.2]).

Fix nn\in\mathbb{N} and let X=S1S1X=S^{1}\vee S^{1} be a wedge of two circles. There exists an 𝔖n\mathfrak{S}_{n}-equivariant isomorphism

gr0WHc3+(2,n)=(Hc(F(X,n))sgn3)𝔖2×𝔖3.\operatorname{gr}_{0}^{W}H_{c}^{3+\ast}(\mathcal{M}_{2,n})=(H_{c}^{\ast}(F(X,n))\otimes\operatorname{sgn}_{3})^{\mathfrak{S}_{2}\times\mathfrak{S}_{3}}. (61)

In this formula, the (𝔖2×𝔖3)(\mathfrak{S}_{2}\times\mathfrak{S}_{3})-action is the following. On sgn3\operatorname{sgn}_{3} it is the sign representation of 𝔖3\mathfrak{S}_{3}, with 𝔖2\mathfrak{S}_{2} acting trivially. On Hc(F(X,n))H_{c}^{\ast}(F(X,n)) it factors through the action of Out(F2)\operatorname{Out}(F_{2}) via a homomorphism 𝔖2×𝔖3Out(F2)\mathfrak{S}_{2}\times\mathfrak{S}_{3}\to\operatorname{Out}(F_{2}). Using Nielsen’s identification Out(F2)GL2()\operatorname{Out}(F_{2})\cong\operatorname{GL}_{2}(\mathbb{Z}) [Nie17], the latter homomorphism is the 22-dimensional representation sgn2std3\operatorname{sgn}_{2}\otimes\operatorname{std}_{3}, where sgn2\operatorname{sgn}_{2} is the sign representation of the 𝔖2\mathfrak{S}_{2} factor and std3=3/(1,1,1)\operatorname{std}_{3}=\mathbb{Z}^{3}/\langle(1,1,1)\rangle is the standard 22-dimensional representation of the 𝔖3\mathfrak{S}_{3} factor.

With this fact at hand we can proceed to prove Proposition 1.10 from the introduction.

Proof of Proposition 1.10.

By [Nie17], abelianization gives Out(F2)GL2()\operatorname{Out}(F_{2})\cong\operatorname{GL}_{2}(\mathbb{Z}). Since Theorem 1.5 shows that the associated graded grHc(F(X,n))\operatorname{gr}H^{*}_{c}(F(X,n)) is a polynomial representation of GL2()\operatorname{GL}_{2}(\mathbb{Z}), and since those are semi-simple, the collision filtration splits canonically so that Hc(F(X,n))grHc(F(X,n))H^{*}_{c}(F(X,n))\cong\operatorname{gr}H^{*}_{c}(F(X,n)) as 𝔖n×Out(F2)\mathfrak{S}_{n}\times\operatorname{Out}(F_{2})-representations.

The dimension of (Hc(F(X,n))sgn3)𝔖2×𝔖3(H_{c}^{\ast}(F(X,n))\otimes\operatorname{sgn}_{3})^{\mathfrak{S}_{2}\times\mathfrak{S}_{3}} can be computed as the scalar product of 𝔖2×𝔖3\mathfrak{S}_{2}\times\mathfrak{S}_{3}-characters

triv,Res𝔖2×𝔖3GL2()Hc(F(X,n))sgn3=sgn3,Res𝔖2×𝔖3GL2()Hc(F(X,n)).\left\langle\operatorname{triv},\ \operatorname{Res}_{\mathfrak{S}_{2}\times\mathfrak{S}_{3}}^{\operatorname{GL}_{2}(\mathbb{Z})}{H_{c}^{\ast}(F(X,n))}\otimes\operatorname{sgn}_{3}\right\rangle=\left\langle\operatorname{sgn}_{3},\ \operatorname{Res}_{\mathfrak{S}_{2}\times\mathfrak{S}_{3}}^{\operatorname{GL}_{2}(\mathbb{Z})}{H_{c}^{\ast}(F(X,n))}\right\rangle.

To remember the 𝔖n\mathfrak{S}_{n}-action one may think of this calculation taking place in the ring of virtual 𝔖n\mathfrak{S}_{n}-representations.

Expand the polynomial functor Hc(F(X,n))H^{*}_{c}(F(X,n)) as in (41),

Hc(F(X,n))λΦ[n,λ]𝕊λ(2)H^{*}_{c}(F(X,n))\cong\bigoplus_{\lambda}\Phi[n,{\lambda}^{\ast}]\boxtimes\mathbb{S}_{{\lambda}}(\mathbb{Q}^{2})

where each Φ[n,λ]\Phi[n,{\lambda}^{\ast}] is an 𝔖n\mathfrak{S}_{n}-representation. Here the conjugate partition λ{\lambda}^{\ast} appears since H~(X)2\tilde{H}^{*}(X)\cong\mathbb{Q}^{2} is concentrated in odd grading. It follows that the above character inner product expands similarly,

sgn3,λRes𝔖2×𝔖3GL2()Φ[n,λ]𝕊λ(2)=λsgn3,Res𝔖2×𝔖3GL2()𝕊λ(2)Φ[n,λ].\begin{split}&\left\langle\operatorname{sgn}_{3},\ \sum_{\lambda}\operatorname{Res}_{\mathfrak{S}_{2}\times\mathfrak{S}_{3}}^{\operatorname{GL}_{2}(\mathbb{Z})}{\Phi[n,{\lambda}^{\ast}]\boxtimes\mathbb{S}_{{\lambda}}(\mathbb{Q}^{2})}\right\rangle\\ &=\sum_{\lambda}\left\langle\operatorname{sgn}_{3},\ \operatorname{Res}_{\mathfrak{S}_{2}\times\mathfrak{S}_{3}}^{\operatorname{GL}_{2}(\mathbb{Z})}{\mathbb{S}_{{\lambda}}(\mathbb{Q}^{2})}\right\rangle\cdot\Phi[n,{\lambda}^{\ast}].\end{split} (62)

This implies the first part of Proposition 1.10, identifying the multiplicity of the 𝔖n\mathfrak{S}_{n}-irreducible χλ\chi_{\lambda} as

rλ=Res𝔖2×𝔖3GL2()𝕊λ(2);sgn3.r_{\lambda}=\left\langle\operatorname{Res}_{\mathfrak{S}_{2}\times\mathfrak{S}_{3}}^{\operatorname{GL}_{2}(\mathbb{Z})}{\mathbb{S}_{\lambda}(\mathbb{Q}^{2})}\ ;\ \operatorname{sgn}_{3}\right\rangle. (63)

The values of the coefficients rλr_{\lambda} are given by the following lemma. ∎

Remark 6.2.

The formula in Proposition 1.10 first came to our knowledge via private communications with Dan Petersen, who derived this formula during ongoing work joint with Orsola Tommasi. Interestingly, in their work the coefficients rλr_{\lambda} are defined as

rλ=dimgr0WHc3(2,𝕍λ)r_{\lambda}=\dim\operatorname{gr}_{0}^{W}H_{c}^{3}(\mathcal{M}_{2},\mathbb{V}_{\lambda})

with 𝕍λ\mathbb{V}_{\lambda} the local system associated to the corresponding irreducible representation of the symplectic group Sp4\operatorname{Sp}_{4}.

Due to the restriction on the coefficient Φ[n,n]\Phi[n,n] given in Proposition 5.5, it follows that our definition of rλr_{\lambda} agrees with theirs. We plan to investigate further this coincidence in future work.

Lemma 6.3.

The coefficients rλr_{\lambda} from Proposition 1.10 are the following. For λ=(a,b)\lambda=(a,b) with ab0a\geq b\geq 0,

r(a,b)={ab6+1if a2b21ab6if a2b200if a2b.r_{(a,b)}=\begin{cases}\left\lfloor{\frac{a-b}{6}}\right\rfloor+1&\text{if }a\equiv_{2}b\equiv_{2}1\\ \left\lfloor{\frac{a-b}{6}}\right\rfloor&\text{if }a\equiv_{2}b\equiv_{2}0\\ 0&\text{if }a\not\equiv_{2}b.\end{cases}
Proof.

The group 𝔖2×𝔖3\mathfrak{S}_{2}\times\mathfrak{S}_{3} is included in GL2()\operatorname{GL}_{2}(\mathbb{Z}) by sending the transposition (12)𝔖2(12)\in\mathfrak{S}_{2} to I2-I_{2}, transpositions in 𝔖3\mathfrak{S}_{3} map to matrices with eigenvalues (1,1)(1,-1) and its 33-cycles map to matrices with eigenvalues (ζ3,ζ31)(\zeta_{3},\zeta_{3}^{-1}) where ζ3\zeta_{3} is a primitive 33rd root of unity.

The character of the Schur representation 𝕊(a,b)(2)\mathbb{S}_{(a,b)}(\mathbb{Q}^{2}) is given by sending a matrix AGL2()A\in\operatorname{GL}_{2}(\mathbb{Z}) with eigenvalues (x,y)(x,y) to the Schur polynomial s(a,b)(x,y)=i=baxiya+bis_{(a,b)}(x,y)=\sum_{i=b}^{a}x^{i}y^{a+b-i} (see [Mac95, §I.3]). One readily checks that transpositions in 𝔖3\mathfrak{S}_{3} acts with trace equal to (1)b(-1)^{b} times the residue of (ab+1)(a-b+1) mod 22, and 33-cycles act with trace (ab+1)(a-b+1) mod 33 (taking 22 mod 33 to have residue 1-1). With these character values the scalar product with sgn3\operatorname{sgn}_{3} is as claimed. ∎

Example 6.4 (Multiplicity of χ(2,1n2)\chi_{(2,1^{n-2})} and χ(n1,1)\chi_{(n-1,1)}).

Let us use calculations of bead representations to see that the 𝔖n\mathfrak{S}_{n}-representation χ(2,1n2)Stdnsgnn\chi_{(2,1^{n-2})}\cong\operatorname{Std}_{n}\otimes\operatorname{sgn}_{n} never occurs in gr0WHcn+2(2,n)\operatorname{gr}^{W}_{0}H^{n+2}_{c}(\mathcal{M}_{2,n}), as conjectured in [BCGY21, Conj. 3.5]. Indeed, in Example 4.18 we explained that Powell–Vespa effectively compute χ(2,1n2)\chi_{(2,1^{n-2})} to never occur in the cohomology Hcn1(F(X,n))H^{n-1}_{c}(F(X,n)). In light of Proposition 1.10, this immediately implies that this irreducible representation does not appear in gr0WHcn+2(2,n)\operatorname{gr}^{W}_{0}H^{n+2}_{c}(\mathcal{M}_{2,n}).

The other half of the conjecture involves the standard representation χ(n1,1)\chi_{(n-1,1)}. After a preprint of the present paper was made public the multiplicity of this representation has been computed by Powell [Pow22]. Combining his result with Proposition 1.10 shows that gr0WHcn+3(2,n)\operatorname{gr}^{W}_{0}H^{n+3}_{c}(\mathcal{M}_{2,n}) contains this representation with the conjectured multiplicity. Since the equivariant Euler characteristic of 2,n\mathcal{M}_{2,n} is known it follows that gr0WHcn+2(2,n)\operatorname{gr}^{W}_{0}H^{n+2}_{c}(\mathcal{M}_{2,n}) also contains this representation with the conjectured multiplicity.

Together with our computations of Φ[n,λ]\Phi[n,\lambda] for n10n\leq 10 below, Proposition 1.10 recovers the 𝔖n\mathfrak{S}_{n}-character of gr0WHc(2,n)\operatorname{gr}_{0}^{W}H_{c}^{*}(\mathcal{M}_{2,n}). For n=11n=11 we have computed the relevant summands of Φ[11,λ]\Phi[11,\lambda], and obtained the result shown below. All our computations agree with those present in [BCGY21].

gr0WHc13(\displaystyle\operatorname{gr}_{0}^{W}H_{c}^{13}( 2,11,)=3χ(9,12)+3χ(8,3)+5χ(8,2,1)+3χ(8,13)+2χ(7,4)+16χ(7,3,1)+\displaystyle\mathcal{M}_{2,11},\mathbb{Q})=3\chi_{(9,1^{2})}+3\chi_{(8,3)}+5\chi_{(8,2,1)}+3\chi_{(8,1^{3})}+2\chi_{(7,4)}+16\chi_{(7,3,1)}+
5χ(7,22)+16χ(7,2,12)+2χ(7,14)+4χ(6,5)+15χ(6,4,1)+23χ(6,3,2)+\displaystyle 5\chi_{(7,2^{2})}+16\chi_{(7,2,1^{2})}+2\chi_{(7,1^{4})}+4\chi_{(6,5)}+15\chi_{(6,4,1)}+23\chi_{(6,3,2)}+
28χ(6,3,12)+24χ(6,22,1)+21χ(6,2,13)+5χ(6,15)+10χ(52,1)+19χ(5,4,2)+\displaystyle 28\chi_{(6,3,1^{2})}+24\chi_{(6,2^{2},1)}+21\chi_{(6,2,1^{3})}+5\chi_{(6,1^{5})}+10\chi_{(5^{2},1)}+19\chi_{(5,4,2)}+
28χ(5,4,12)+21χ(5,32)+50χ(5,3,2,1)+28χ(5,3,13)+13χ(5,23)+38χ(5,22,12)+\displaystyle 28\chi_{(5,4,1^{2})}+21\chi_{(5,3^{2})}+50\chi_{(5,3,2,1)}+28\chi_{(5,3,1^{3})}+13\chi_{(5,2^{3})}+38\chi_{(5,2^{2},1^{2})}+
17χ(5,2,14)+7χ(5,16)+8χ(42,3)+29χ(42,2,1)+20χ(42,13)+25χ(4,32,1)+\displaystyle 17\chi_{(5,2,1^{4})}+7\chi_{(5,1^{6})}+8\chi_{(4^{2},3)}+29\chi_{(4^{2},2,1)}+20\chi_{(4^{2},1^{3})}+25\chi_{(4,3^{2},1)}+
28χ(4,3,22)+48χ(4,3,2,12)+22χ(4,3,14)+22χ(4,23,1)+25χ(4,22,13)+\displaystyle 28\chi_{(4,3,2^{2})}+48\chi_{(4,3,2,1^{2})}+22\chi_{(4,3,1^{4})}+22\chi_{(4,2^{3},1)}+25\chi_{(4,2^{2},1^{3})}+
11χ(4,2,15)+2χ(4,17)+13χ(33,2)+8χ(33,12)+22χ(32,22,1)+20χ(32,2,13)+\displaystyle 11\chi_{(4,2,1^{5})}+2\chi_{(4,1^{7})}+13\chi_{(3^{3},2)}+8\chi_{(3^{3},1^{2})}+22\chi_{(3^{2},2^{2},1)}+20\chi_{(3^{2},2,1^{3})}+
11χ(32,15)+4χ(3,24)+15χ(3,23,12)+8χ(3,22,14)+6χ(3,2,16)+3χ(25,1)+\displaystyle 11\chi_{(3^{2},1^{5})}+4\chi_{(3,2^{4})}+15\chi_{(3,2^{3},1^{2})}+8\chi_{(3,2^{2},1^{4})}+6\chi_{(3,2,1^{6})}+3\chi_{(2^{5},1)}+
4χ(24,13)+2χ(23,15)+2χ(22,17)+χ(111),\displaystyle 4\chi_{(2^{4},1^{3})}+2\chi_{(2^{3},1^{5})}+2\chi_{(2^{2},1^{7})}+\chi_{(1^{11})},

which has dimension 850732850732.

For numbers nn larger still, the super-exponential multiplicity of every Schur functor discussed in Example 5.24 implies similar growth in gr0WHc(2,n)\operatorname{gr}^{W}_{0}H^{*}_{c}(\mathcal{M}_{2,n})111111The 𝔖n\mathfrak{S}_{n}-equivariant Euler characteristic of these representations is computed in [CFGP19], and gives a lower bound on multiplicity of all χλ\chi_{\lambda} in cohomology. Our lower bound is orthogonal, and shows the codimension 11 cohomology is much larger than can be deduced from [CFGP19].. The reader can find our equivariant lower bound for gr0WHc(2,n)\operatorname{gr}_{0}^{W}H_{c}^{*}(\mathcal{M}_{2,n}) presented on this webpage121212https://louishainaut.github.io/GH-ConfSpace/. Forgetting the 𝔖n\mathfrak{S}_{n}-action, a lower bound on the dimension of the cohomology gr0WHcn+2(2,n)\operatorname{gr}^{W}_{0}H^{n+2}_{c}(\mathcal{M}_{2,n}) for all n17n\leq 17 are listed in Table 2.

nn 0 11 22 33 44 55 66 77 88 99 1010
dim\dim 0 0 0 0 11 55 2626 155155 10661066 86668666 8101281012
nn 1111 1212 1313 1414
dim\dim 850 732850\,732 7 920 155\geq 7\,920\,155 94 325 925\geq 94\,325\,925 1 220 494 146\geq 1\,220\,494\,146
nn 1515 1616 1717
dim\dim 17 048 375 436\geq 17\,048\,375\,436 255 669 776 040\geq 255\,669\,776\,040 4 096 729 778 379\geq 4\,096\,729\,778\,379
Table 2: Dimension of gr0WHcn+2(2,n,)\operatorname{gr}_{0}^{W}H_{c}^{n+2}(\mathcal{M}_{2,n},\mathbb{Q})

6.2 Patterns and conjectures

We start this section by proving the fourth statement of Proposition 5.5,

Proposition 6.5.

For a wedge of circles X=i=1gS1X=\bigvee_{i=1}^{g}S^{1},

gr0Hcn1(F(X,n))=χ(1n)Symn1(g).\operatorname{gr}_{0}H_{c}^{n-1}(F(X,n))=\chi_{(1^{n})}\boxtimes\operatorname{Sym}^{n-1}(\mathbb{Q}^{g}).

Equivalently, the coefficient Φ0[n,n1]\Phi^{0}[n,n-1] of the polynomial decomposition is

Φ0[n,n1]=χ(1n)χ(1n1).\Phi^{0}[n,n-1]=\chi_{(1^{n})}\boxtimes\chi_{{(1^{n-1})}}.
Proof.

From Proposition 5.16 the equivariant multiplicity of the alternating power Λn1\Lambda^{n-1} in Ψ0(𝐧,)\Psi^{0}(\mathbf{n},-) is

Φ0[n,(1n1)]=H0(F(3,n1))sgnn=sgnn,\Phi^{0}[n,{(1^{n-1})}]=H_{0}(F(\mathbb{R}^{3},n-1))\otimes\operatorname{sgn}_{n}=\operatorname{sgn}_{n},

so Φ0[n,n1]\Phi^{0}[n,n-1] is at least as large as indicated, and it remains to prove that it does not contain any other terms.

For this we look closely at the collision spectral sequence. Once again, it is enough to prove the statement when X=i=1gS1X=\bigvee_{i=1}^{g}{S^{1}} is a wedge of sufficiently many circles. In fact gn1g\geq n-1 is enough, but we may as well consider all gg’s simultaneously. The polynomial representation Φ0[n,n1]𝔖n1g[1]n1\Phi^{0}[n,n-1]\otimes_{\mathfrak{S}_{n-1}}{\mathbb{Q}^{g}[-1]}^{\otimes n-1} is by definition the kernel of the differential d1:E10,n1E11,n1d_{1}\colon E_{1}^{0,n-1}\to E_{1}^{1,n-1}. The multiplicity of alternating powers implies that this kernel contains at least one copy of sgnnSymn1(g)[1n]\operatorname{sgn}_{n}\boxtimes\operatorname{Sym}^{n-1}(\mathbb{Q}^{g})[1-n], which has dimension (n+g2n1)\binom{n+g-2}{n-1}. We claim that the kernel of this differential cannot have larger dimension.

A basis for E10,n1=Λn(H(X)SLie(1))n1E_{1}^{0,n-1}=\Lambda^{n}\left(H^{*}(X)\otimes\operatorname{S}\operatorname{Lie}(1)\right)^{n-1} is given by tuples

(x1,αi1),,(xk1,αik1),(xk,1),(xk+1,αik+1),,(xn,αin)\langle(x_{1},\alpha_{i_{1}}),\ldots,(x_{k-1},\alpha_{i_{k-1}}),(x_{k},1),(x_{k+1},\alpha_{i_{k+1}}),\ldots,(x_{n},\alpha_{i_{n}})\rangle

where the symbols xix_{i} are copies of the ‘Lie variable’ in SLie(1)x\operatorname{S}\operatorname{Lie}(1)\cong\mathbb{Q}x, and the vectors α1,,αgH~1(X)\alpha_{1},\ldots,\alpha_{g}\in\tilde{H}^{1}(X) form a basis. Here 11 refers to the unit 1H0(X)1\in H^{0}(X), and the indices are such that 1kn1\leq k\leq n along with 1i1,,ik^,,ing1\leq i_{1},\ldots,\widehat{i_{k}},\ldots,i_{n}\leq g. We denote this basis element by (i1,,ik^,,in)\mathcal{B}_{(i_{1},\ldots,\widehat{i_{k}},\ldots,i_{n})}.

For E11,n1(n2)(H(X)SLie(2))Λn2(H(X)SLie(1))n2E_{1}^{1,n-1}\cong\bigoplus_{\binom{n}{2}}(H^{*}(X)\otimes\operatorname{S}\operatorname{Lie}(2))\bigotimes\Lambda^{n-2}\left(H^{*}(X)\otimes\operatorname{S}\operatorname{Lie}(1)\right)^{n-2} we work with the basis of tuples

([xj,xk],αi0),(x1,αi1),,(xj,αij)^,,(xk,αik)^,,(xn,αin)\langle([x_{j},x_{k}],\alpha_{i_{0}}),(x_{1},\alpha_{i_{1}}),\ldots,\widehat{(x_{j},\alpha_{i_{j}})},\ldots,\widehat{(x_{k},\alpha_{i_{k}})},\ldots,(x_{n},\alpha_{i_{n}})\rangle

for every pair {j,k}\{j,k\} and sequence of indices (i0,i1,,ij^,,ik^,,in)(i_{0},i_{1},\ldots,\widehat{i_{j}},\ldots,\widehat{i_{k}},\ldots,i_{n}). Note that now all terms in the tuple contain a cohomology class αiH~1(X)\alpha_{i}\in\tilde{H}^{1}(X). We denote this basis element by 𝒞(i0,i1,,i^j,,i^k,,in)\mathcal{C}_{(i_{0},i_{1},\ldots,\widehat{i}_{j},\ldots,\widehat{i}_{k},\ldots,i_{n})}.

Now, the differential d1d_{1} acts on our basis by

d1:(i1,,ik^,,in)jk±𝒞(ij,i1,,i^j,,i^k,,n)d_{1}:\mathcal{B}_{(i_{1},\ldots,\widehat{i_{k}},\ldots,i_{n})}\longmapsto\sum_{j\neq k}\pm\mathcal{C}_{(i_{j},i_{1},\ldots,\widehat{i}_{j},\ldots,\widehat{i}_{k},\ldots,n)} (64)

and in particular all resulting basis elements appear with coefficient ±1\pm 1. Let us consider the dual problem: which 𝐢\mathcal{B}_{\mathbf{i}}’s are sent to a linear combination containing a nonzero multiple of a particular c=𝒞(i0,,i^j,,i^k,,in)c=\mathcal{C}_{(i_{0},\ldots,\widehat{i}_{j},\ldots,\widehat{i}_{k},\ldots,i_{n})}?

From the description of d1d_{1} there are exactly two basis elements with this property: b1=(i1,,i0,,ik^,,in)b_{1}=\mathcal{B}_{(i_{1},\ldots,i_{0},\ldots,\widehat{i_{k}},\ldots,i_{n})} and b2=(i1,,i^j,,i0,,in)b_{2}=\mathcal{B}_{(i_{1},\ldots,\widehat{i}_{j},\ldots,i_{0},\ldots,i_{n})}, with i0i_{0} inserted in the jj-th, resp. kk-th, empty slot. It follows that every element of ker(d1)\ker(d_{1}) that contains a nontrivial multiple of b1b_{1} must also contain one of b2b_{2}, and the coefficient of one uniquely determines the coefficient of the other.

Define an equivalence relation of the 𝐢\mathcal{B}_{\mathbf{i}}’s, generated by the relation that b1b2b_{1}\sim b_{2} if their indexing tuples differ by moving one iji_{j} of b1b_{1} to the empty slot of b2b_{2}. Then the discussion in the previous paragraph shows that ker(d1)\ker(d_{1}) is spanned by \sim-equivalence classes, and it remains to count these classes.

It is straightforward to see that equivalence classes are in bijection with multisets i1,,in1\langle i_{1},\ldots,i_{n-1}\rangle with 1ijg1\leq i_{j}\leq g for all jj. Thus there are exactly (n+g2n1)\binom{n+g-2}{n-1} many equivalence classes, and the kernel has at most this dimension. This is what we wanted to show. ∎

Remark 6.6.

Using the Euler characteristic as in §5.4, one can obtain from the previous proposition a plethystic description of gr1Hcn(F(X,n))\operatorname{gr}_{1}H^{n}_{c}(F(X,n)), equivalently Φ1[n,n1]\Phi^{1}[n,n-1], identical to the one in [PV18, Corollary 3].

Calculations for small values of nn also suggest the following pattern.

Conjecture 6.7.

For n4n\geq 4 and X=gS1X=\bigvee_{g}S^{1} a wedge of circles,

gr1Hcn1(F(X,n))=(χ(3,1n3)Symn2(g))(χ(n)Λn2(g)).\operatorname{gr}_{1}H_{c}^{n-1}(F(X,n))=\big{(}\chi_{(3,1^{n-3})}\boxtimes\operatorname{Sym}^{n-2}(\mathbb{Q}^{g})\big{)}\oplus\big{(}\chi_{(n)}\boxtimes\Lambda^{n-2}(\mathbb{Q}^{g})\big{)}.

Equivalently, the coefficients of Φ1[n,n2]\Phi^{1}[n,n-2] are given by

Φ1[n,n2]=(χ(3,1n3)χ(1n2))(χ(n)χ(n2)).\Phi^{1}[n,n-2]=(\chi_{(3,1^{n-3})}\boxtimes\chi_{(1^{n-2})})\oplus(\chi_{(n)}\boxtimes\chi_{(n-2)}). (65)

Up to this point we mainly discussed the 𝔖n\mathfrak{S}_{n}-equivariant multiplicity of individual Schur functors in grHc(F(X,n))\operatorname{gr}H^{*}_{c}(F(X,n)), equivalently the coefficients Φ[,μ]\Phi[-,\mu]. Let us now shift perspective and instead consider for fixed 𝔖n\mathfrak{S}_{n}-irreducible χλ\chi_{\lambda} the corresponding isotypic component as a sum of Schur functors. These are equivalently given by the coefficient Φ[λ,]:=m=0|λ|Φ[λ,m]\Phi[\lambda,-]:=\oplus_{m=0}^{|\lambda|}\Phi[\lambda,m].

Through our calculations we have identified several patterns in these coefficients – see below. We present our conjectural formulas only for multiplicities in codimension 1 cohomology of F(S1,n)F(\vee S^{1},n), but using the spectral sequence to compute the Euler characteristic, one can readily compute the corresponding multiplicity in codimension 0 as well.

In some of the conjectural patterns below we use the notation

Mn=2a+b=n𝕊(a,1b), a sum of hook shapes, M_{n}=\sum_{2a+b=n}{\mathbb{S}_{{(a,1^{b})}}},\text{ a sum of hook shapes, }

where the sum is over a1a\geq 1 and b0b\geq 0. The following can be deduced from [PV18, Examples 3-4 and Corollary 19.8] using the dictionary in §4.4.

Proposition 6.8.

For partition λn\lambda\vdash n, the equivariant multiplicity of χλ\chi_{\lambda} in grHcn1(F(X,n))\operatorname{gr}H^{n-1}_{c}(F(X,n)), equivalently the bead representation UλIIU_{\lambda}^{II} from [TW19, §2.5], is the following polynomial functor.

  • For the trivial representation λ=(n)\lambda=(n), it is Mn1M_{n-1}.

  • For the sign representation λ=(1n)\lambda=(1^{n}), it is 𝕊(n1)\mathbb{S}_{(n-1)}.

  • For λ=(2,1n2)\lambda=(2,1^{n-2}) it is 0.

We observed the following further patterns, verified computationally for up to n11n\leq 11 and compliant with the lower bound in §5.4. Any multiplicity involving 𝕊(k)\mathbb{S}_{{(k)}^{\ast}} and 𝕊(1k)\mathbb{S}_{{(1^{k})}^{\ast}} is completely determined by Theorem 1.8.

Conjecture 6.9.

The equivariant multiplicity of χλ\chi_{\lambda} in grHcn1(F(X,n))\operatorname{gr}H_{c}^{n-1}(F(X,n)) are:

  • For λ=(n1,1)\lambda=(n-1,1) it is {𝕊(n/2)if n is even0if n is odd\begin{cases}\mathbb{S}_{{(n/2)}}&\text{if }n\text{ is even}\\ 0&\text{if }n\text{ is odd}\end{cases} (proved in Prop 6.12 below).

  • For λ=(n2,2)\lambda=(n-2,2) it is the sum k=0n41Mn24k+k=0n74Mn54k\sum_{k=0}^{\lfloor{\frac{n}{4}}\rfloor-1}{M_{n-2-4k}}+\sum_{k=0}^{\lfloor{\frac{n-7}{4}}\rfloor}{M_{n-5-4k}}

    +{k=0n42𝕊(n422k)if n is evenk=0n54𝕊(n122k)if n is odd+\left\{\begin{array}[]{ll}\sum_{k=0}^{\lfloor{\frac{n}{4}}\rfloor-2}\mathbb{S}_{{\left(\frac{n-4}{2}-2k\right)}}&\text{if }n\text{ is even}\\ \sum_{k=0}^{\lfloor{\frac{n-5}{4}}\rfloor}\mathbb{S}_{{\left(\frac{n-1}{2}-2k\right)}}&\text{if }n\text{ is odd}\end{array}\right.
  • For λ=(n2,1,1)\lambda=(n-2,1,1) it is the sum k=0n54Mn34k+k=0n64Mn44k\sum_{k=0}^{\lfloor{\frac{n-5}{4}}\rfloor}{M_{n-3-4k}}+\sum_{k=0}^{\lfloor{\frac{n-6}{4}}\rfloor}{M_{n-4-4k}}

    +{k=0n64𝕊(n222k)if n is evenk=0n34𝕊(n+122k)if n is odd+\left\{\begin{array}[]{ll}\sum_{k=0}^{\lfloor{\frac{n-6}{4}}\rfloor}\mathbb{S}_{{\left(\frac{n-2}{2}-2k\right)}}&\text{if }n\text{ is even}\\ \sum_{k=0}^{\lfloor{\frac{n-3}{4}}\rfloor}\mathbb{S}_{{\left(\frac{n+1}{2}-2k\right)}}&\text{if }n\text{ is odd}\end{array}\right.
  • For λ=(n3,3)\lambda={(n-3,3)}^{\ast} it is k=0n6k+33𝕊(n4k)+an𝕊(1,1)+bn𝕊(1)\sum_{k=0}^{n-6}{\lfloor{\frac{k+3}{3}}\rfloor\mathbb{S}_{{(n-4-k)}}}+a_{n}\mathbb{S}_{{(1,1)}}+b_{n}\mathbb{S}_{{(1)}}

  • For λ=(n3,2,1)\lambda={(n-3,2,1)}^{\ast} it is k=0n52k+33𝕊(n3k)+n23𝕊(1,1)+n43𝕊(1)\sum_{k=0}^{n-5}{\left\lfloor{\frac{2k+3}{3}}\right\rfloor\mathbb{S}_{{(n-3-k)}}}+\left\lfloor{\frac{n-2}{3}}\right\rfloor\mathbb{S}_{{(1,1)}}+\left\lfloor{\frac{n-4}{3}}\right\rfloor\mathbb{S}_{{(1)}}

  • For λ=(n3,1,1,1)\lambda={(n-3,1,1,1)}^{\ast} it is k=0n6k+33𝕊(n4k)+cn𝕊(1,1)+dn𝕊(1)\sum_{k=0}^{n-6}{\left\lfloor{\frac{k+3}{3}}\right\rfloor\mathbb{S}_{{(n-4-k)}}}+c_{n}\mathbb{S}_{{(1,1)}}+d_{n}\mathbb{S}_{{(1)}}

  • For λ=(n2,2)\lambda={(n-2,2)}^{\ast} it is k=0n/22𝕊(n32k)\sum_{k=0}^{\lfloor{n/2}\rfloor-2}{\mathbb{S}_{{(n-3-2k)}}}

  • For λ=(n2,1,1)\lambda={(n-2,1,1)}^{\ast} it is k=0(n3)/2𝕊(n22k)\sum_{k=0}^{\lfloor{(n-3)/2}\rfloor}{\mathbb{S}_{{(n-2-2k)}}}

where the numbers an,bn,cna_{n},b_{n},c_{n} and dnd_{n} are n/6\lfloor{n/6}\rfloor plus 66-periodic functions of nn. Explicitly,

an=n/6\displaystyle a_{n}=\lfloor{n/6}\rfloor +\displaystyle+ (1,0,1,0,0,0)n\displaystyle(-1,0,-1,0,0,0)_{n}
bn=n/6\displaystyle b_{n}=\lfloor{n/6}\rfloor +\displaystyle+ (1,1,1,0,1,0)n\displaystyle(-1,-1,-1,0,-1,0)_{n}
cn=n/6\displaystyle c_{n}=\lfloor{n/6}\rfloor +\displaystyle+ (0,0,0,0,1,0)n\displaystyle(0,0,0,0,1,0)_{n}
dn=n/6\displaystyle d_{n}=\lfloor{n/6}\rfloor +\displaystyle+ (0,1,0,0,0,0)n\displaystyle(0,-1,0,0,0,0)_{n}

where ()n(-)_{n} returns the entry in position residue of nn modulo 66 (starting at residue 0).

Remark 6.10.

Interestingly, the complicated patterns for λ=(n2,2)\lambda=(n-2,2) and (n2,1,1)(n-2,1,1) become far simpler when summed together; and the same is true for their conjugate partitions. For the partitions λ=(n3,3)\lambda={(n-3,3)}^{\ast} and (n3,1,1,1){(n-3,1,1,1)}^{\ast} it is now their difference that has a significantly simpler form: one can observe that cnanc_{n}-a_{n} and dnbnd_{n}-b_{n} both take the value 11 if nn is even and 0 if nn is odd. We have no guess as to why this should be the case.

Remark 6.11.

After a first version of this paper was made public, Powell [Pow22] computed the composition factors of χ(n1,1)\chi_{(n-1,1)} in grHcn(F(X,n))\operatorname{gr}H_{c}^{n}(F(X,n)), related to the first case of Conjecture 6.9 via Euler characteristic. That said, verifying that the conjecture follows from Powell’s work is challenging, since calculating the Euler characteristic involves plethysm coefficients whose determination is still an open problem. In the next proposition we prove this case of the conjecture by combining several ideas from this paper.

Proposition 6.12.

The equivariant multiplicity of χ(n1,1)\chi_{(n-1,1)} in codimension 11 cohomology grHcn1(F(X,n))\operatorname{gr}H_{c}^{n-1}(F(X,n)) is 𝕊(n/2)\mathbb{S}_{{(n/2)}} when nn is even, and 0 when nn is odd.

Proof.

Powell [Pow22, Theorem 3] provides the equivariant multiplicity of χ(n1,1)\chi_{(n-1,1)} in codimension 0 cohomology grHcn(F(X,n))\operatorname{gr}H_{c}^{n}(F(X,n)). To see that the equivariant multiplicity in codimension 11 is at least as large as we claim, we note that when nn is even, our Proposition 5.16 shows that the 𝔖n\mathfrak{S}_{n}-equivariant multiplicity of the Schur functor 𝕊(n/2)\mathbb{S}_{(n/2)} in codimension 11 is the same as that of 𝕊(n/2+1)\mathbb{S}_{(n/2+1)} in codimension 0, and Powell’s calculation shows the representation χn1,1𝕊(n/2+1)\chi_{n-1,1}\boxtimes\mathbb{S}_{(n/2+1)} indeed occurs in codimension 0.

To show that the multiplicity is no larger than claimed, it is therefore enough to bound its dimension. The (non-equivariant) Euler characteristic of the χ(n1,1)\chi_{(n-1,1)}-multiplicity space can be read-off from the 2-step complex (35), and takes the value (1)n(n1)(n+g2g2)(-1)^{n}(n-1)\binom{n+g-2}{g-2}. One can then compute the non-equivariant multiplicity in codimension 0 provided by Powell, and verify that the multiplicity in codimension 11 has the dimension we claim.∎

Using Proposition 1.10, Conjecture 6.9 leads to the following conjectural irreducible multiplicities in gr0WHcn+2(2,n,)\operatorname{gr}_{0}^{W}H_{c}^{n+2}(\mathcal{M}_{2,n},\mathbb{Q}).

Conjecture 6.13.

The irreducible representation χλ\chi_{\lambda} appears in gr0WHcn+2(2,n,)\operatorname{gr}_{0}^{W}H_{c}^{n+2}(\mathcal{M}_{2,n},\mathbb{Q}) with the following multiplicity:

  • For λ=(n2,2)\lambda=(n-2,2) it is {k2k+13if n=4k3k2k+46if n=4k+1k(k1)6if n=4k+20if n=4k+3.\begin{cases}\lfloor\frac{k^{2}-k+1}{3}\rfloor&\text{if }n=4k\\ \lfloor\frac{3k^{2}-k+4}{6}\rfloor&\text{if }n=4k+1\\ \lfloor\frac{k(k-1)}{6}\rfloor&\text{if }n=4k+2\\ 0&\text{if }n=4k+3\\ \end{cases}.

  • For λ=(n2,1,1)\lambda=(n-2,1,1) it is {(k1)(k2)6if n=4k0if n=4k+1k2+k+13if n=4k+23k2+3k+26if n=4k+3.\begin{cases}\lfloor\frac{(k-1)(k-2)}{6}\rfloor&\text{if }n=4k\\ 0&\text{if }n=4k+1\\ \lfloor\frac{k^{2}+k+1}{3}\rfloor&\text{if }n=4k+2\\ \lfloor\frac{3k^{2}+3k+2}{6}\rfloor&\text{if }n=4k+3\\ \end{cases}.

  • For λ=(n2,2)\lambda={(n-2,2)}^{\ast} it is {0if n=2k(k2)(k1)6if n=2k+1.\begin{cases}0&\text{if }n=2k\\ \lfloor\frac{(k-2)(k-1)}{6}\rfloor&\text{if }n=2k+1\\ \end{cases}.

  • For λ=(n2,1,1)\lambda={(n-2,1,1)}^{\ast} it is {(k2)(k1)6if n=2k0if n=2k+1.\begin{cases}\lfloor\frac{(k-2)(k-1)}{6}\rfloor&\text{if }n=2k\\ 0&\text{if }n=2k+1\\ \end{cases}.

The multiplicities in Conjectures 6.9 and 6.13 have nice presentations as generating functions. These can be found in Appendix B.

Appendix A Full decomposition for 1010 particles

We explain here how to combine insights from the different presentations of §4 to compute the full cohomology of F(X,10)F(X,10) for XX any wedge of circles. Conceptually, the 22-term complex of §4.3 gives access to individual 𝔖n\mathfrak{S}_{n}-isotypic components, most effective for Schur functors 𝕊λ\mathbb{S}_{\lambda} for λ\lambda with few parts. On the other hand, the Chevalley–Eilenberg complex of §4.1 accesses individual GL\operatorname{GL}-isotypic components, most effective for 𝕊λ\mathbb{S}_{\lambda} of large total degree |λ||\lambda|. Between these two, one can uniquely determine the complete representation.

To illustrate our approach we focus on the multiplicity of a few specific 𝔖10\mathfrak{S}_{10}-irreducibles in the bottom cohomology H9(F(X,10))H^{9}(F(X,10)). Note that by the bound on polynomial degree in Proposition 5.5 and since the highest degree term Φ0[10,9]\Phi^{0}[10,9] is also computed there, the only remaining multiplicities that need to be computed are those of Schur functors of degree 8\leq 8. In the following we will show how to compute the equivariant multiplicity of the 𝔖10\mathfrak{S}_{10}-representations χ(8,2)\chi_{(8,2)}, χ(6,3,1)\chi_{(6,3,1)}, χ(42,2)\chi_{(4^{2},2)} and χ(3,2,15)\chi_{(3,2,1^{5})}.

Step 1: Lower bound

First compute the lower bounds for these 𝔖10\mathfrak{S}_{10}-irreducible as in §5.4, corrected with the knowledge of the exact multiplicity of the symmetric and alternating powers 𝕊(k)\mathbb{S}_{(k)} and 𝕊(1k)\mathbb{S}_{(1^{k})} from §5.3:

𝔖10\mathfrak{S}_{10}-irrep GL\operatorname{GL}-equivariant multiplicities
(8,2)(8,2) 𝕊(4)+𝕊(3)+𝕊(2,1)+𝕊(2)+𝕊(17)+𝕊(14)+𝕊(13)\geq\mathbb{S}_{(4)}+\mathbb{S}_{(3)}+\mathbb{S}_{(2,1)}+\mathbb{S}_{(2)}+\mathbb{S}_{(1^{7})}+\mathbb{S}_{(1^{4})}+\mathbb{S}_{(1^{3})}
(6,3,1)(6,3,1) 𝕊(5)+6𝕊(4)+11𝕊(3)+2𝕊(2,1)+10𝕊(2)+𝕊(16)+2𝕊(15)+3𝕊(14)+5𝕊(13)+7𝕊(12)+4𝕊(1)\geq\mathbb{S}_{(5)}+6\mathbb{S}_{(4)}+11\mathbb{S}_{(3)}+2\mathbb{S}_{(2,1)}+10\mathbb{S}_{(2)}+\mathbb{S}_{(1^{6})}+2\mathbb{S}_{(1^{5})}+3\mathbb{S}_{(1^{4})}+5\mathbb{S}_{(1^{3})}+7\mathbb{S}_{(1^{2})}+4\mathbb{S}_{(1)}
(42,2)(4^{2},2) 𝕊(5)+6𝕊(4)+6𝕊(3)+𝕊(2,1)+10𝕊(2)+2𝕊(14)+6𝕊(13)+5𝕊(12)+𝕊(1)\geq\mathbb{S}_{(5)}+6\mathbb{S}_{(4)}+6\mathbb{S}_{(3)}+\mathbb{S}_{(2,1)}+10\mathbb{S}_{(2)}+2\mathbb{S}_{(1^{4})}+6\mathbb{S}_{(1^{3})}+5\mathbb{S}_{(1^{2})}+\mathbb{S}_{(1)}
(3,2,15)(3,2,1^{5}) 𝕊(7)+𝕊(6)+2𝕊(5)+3𝕊(4)+3𝕊(3)+4𝕊(2)+2𝕊(12)+2𝕊(1)\geq\mathbb{S}_{(7)}+\mathbb{S}_{(6)}+2\mathbb{S}_{(5)}+3\mathbb{S}_{(4)}+3\mathbb{S}_{(3)}+4\mathbb{S}_{(2)}+2\mathbb{S}_{(1^{2})}+2\mathbb{S}_{(1)}
Table 3: Lower bounds for n=10n=10

We already know the multiplicity of all symmetric powers 𝕊(m)\mathbb{S}_{(m)}, so we proceed to determine the multiplicities of Schur functors with two parts 𝕊(a,b)\mathbb{S}_{(a,b)}.

Step 2: Traces on Schurs with two parts

To gain understanding of the End(2)\operatorname{End}(\mathbb{Z}^{2})-action on each of the above multiplicity spaces, we use the 2-step complex from 4.3 with genus g=2g=2 to compute its character. Start by computing the trace of the action induced by specific diagonal matrices on each multiplicity space: diag(1,1),diag(1,2)diag(1,1),diag(1,2) and diag(2,2)diag(2,2) – these calculations are not too demanding as they only involve one or two doubling maps on S1S1S^{1}\vee S^{1}.

Once these traces are obtained, we subtract from them the traces of the Schur functors appearing in the lower bound. We thus obtain the traces of terms missing from our lower bound. In our example, the traces corresponding to the lower bounds, as well as the true computed traces of our 33 diagonal matrices and the difference between them are as in Table 4

𝔖10\mathfrak{S}_{10}-irreducible Lower bound Computed traces Difference
(8,2)(8,2) [14,59,140][14,59,140] [14,59,140][14,59,140] [0,0,0][0,0,0]
(6,3,1)(6,3,1) [129,522,1220][129,522,1220] [145,580,1396][145,580,1396] [16,58,176][16,58,176]
(42,2)(4^{2},2) [99,428,1024][99,428,1024] [109,464,1136][109,464,1136] [10,36,112][10,36,112]
(3,2,15)(3,2,1^{5}) [72,684,2256][72,684,2256] [72,684,2256][72,684,2256] [0,0,0][0,0,0]
Table 4: Traces for the matrices diag(1,1), diag(1,2) and diag(2,2)

One thus immediately sees that for the first and last rows the lower bound is sharp, i.e. it accounts for all Schur functors 𝕊(a,b)\mathbb{S}_{(a,b)}. The remaining rows are missing a 1616-dimensional and a 1010-dimensional representation respectively, with known character values at diag(1,2)diag(1,2) and diag(2,2)diag(2,2).

The problem thus reduces to finding a sum of Schur functors that produces the traces seen in the last column of Table 4. Note that while the complete traces are rather large, their difference from our lower bound is significantly smaller. Moreover, as stated at the outset of this calculation, only Schur functors of degree 8\leq 8 may appear. The corresponding traces on these Schur functors are given in Table 5.

Schur functor Traces Schur functor Traces
(2,1)(2,1) [2,6,16][2,6,16] (3,1)(3,1) [3,14,48][3,14,48]
(22)(2^{2}) [1,4,16][1,4,16] (4,1)(4,1) [4,30,128][4,30,128]
(3,2)(3,2) [2,12,64][2,12,64] (5,1)(5,1) [5,62,320][5,62,320]
(4,2)(4,2) [3,28,192][3,28,192] (32)(3^{2}) [1,8,64][1,8,64]
(6,1)(6,1) [6,126,768][6,126,768] (5,2)(5,2) [4,60,512][4,60,512]
(4,3)(4,3) [2,24,256][2,24,256] (7,1)(7,1) [7,254,1792][7,254,1792]
(6,2)(6,2) [5,124,1280][5,124,1280] (5,3)(5,3) [3,56,768][3,56,768]
(42)(4^{2}) [1,16,256][1,16,256]
Table 5: Traces of Schur functors at diag(1,1), diag(2,1) and diag(2,2)

It is useful to note that s(a,b)(2,2)=2a+bs(a,b)(1,1)s_{(a,b)}(2,2)=2^{a+b}s_{(a,b)}(1,1) for any Schur polynomial s(a,b)s_{(a,b)}, i.e. the third trace in the table is always 2a+b2^{a+b} larger than the first. Therefore since 𝕊(2,1)\mathbb{S}_{(2,1)} is the only Schur functor in our table for which 2a+b<162^{a+b}<16, any sum of other Schur functors must have its third trace at least 1616 times larger than than the first. Since the missing Schur functors for χ(6,3,1)\chi_{(6,3,1)} and χ(42,2)\chi_{(4^{2},2)} have their third trace <16<16 times their first, one immediately concludes that 𝕊(2,1)\mathbb{S}_{(2,1)} occurs several times.

Let us therefore account for these copies of 𝕊(2,1)\mathbb{S}_{(2,1)} and remove them from the trace until the third column is 16\geq 16 times the first. For example, from the traces [10,36,112][10,36,112] corresponding to χ(42,2)\chi_{(4^{2},2)} we subtract m×[2,6,16]m\times[2,6,16] with mm such that 16(102m)11216m16\cdot(10-2m)\leq 112-16m. That is, m3m\geq 3, leaving traces [4,18,64][4,18,64] to be accounted for. In general the 𝕊(2,1)\mathbb{S}_{(2,1)}-multiplicity is c1c3/16\geq c_{1}-c_{3}/16 where cic_{i} is the ii-th trace in our table.

At this point the remaining Schur functors can be uniquely determined. The Schur functors 𝕊(a,b)\mathbb{S}_{(a,b)} that need to be added can be found in Table 6; the interested reader is invited to verify that no other combination of Schur functors gives the desired traces.

𝔖10\mathfrak{S}_{10}-irreducible Missing Schur functors
(8,2)(8,2) 0
(6,3,1)(6,3,1) 5𝕊(2,1)+2𝕊(3,1)5\mathbb{S}_{(2,1)}+2\mathbb{S}_{(3,1)}
(42,2)(4^{2},2) 3𝕊(2,1)+𝕊(3,1)+𝕊(2,2)3\mathbb{S}_{(2,1)}+\mathbb{S}_{(3,1)}+\mathbb{S}_{(2,2)}
(3,2,15)(3,2,1^{5}) 0
Table 6: Corrections to the lower bounds in genus 22

We therefore add these Schur functors to our lower bound estimate and proceed to find Schur functors with 33 or more parts.

Step 3: Ranks in genus 3 and 4

One could attempt the same game in higher genus, i.e. finding multiplicities of 𝕊(a,b,c)\mathbb{S}_{(a,b,c)} and so on, but it turns out that one can (almost always) make due with only the nonequivariant ranks in genus 33 and 44. For example, Table 7 shows the ranks in genus 33 and 44 of the revised lower bound compared with the true rank computed using the 22-step complex in §4.3.

𝔖10\mathfrak{S}_{10}-irreducible Lower bound Actual ranks Difference
(8,2)(8,2) [40,90][40,90] [46,126][46,126] [6,36][6,36]
(6,3,1)(6,3,1) [405,897][405,897] [411,931][411,931] [6,34][6,34]
(42,2)(4^{2},2) [308,691][308,691] [308,691][308,691] [0,0][0,0]
(3,2,15)(3,2,1^{5}) [217,541][217,541] [217,541][217,541] [0,0][0,0]
Table 7: Ranks in genus 3 and 4

From Table 7 one sees that only rather small Schur functors may appear. Table 8 lists the ranks of all Schur functors of degree 8\leq 8 with 33 or 44 parts. For most Specht modules of 𝔖10\mathfrak{S}_{10} this table is already sufficient for uniquely determining the multiplicity of every Schur functor with 33 or 44 parts.

Schur functor Ranks in genus 3,43,4 Schur functor Ranks in genus 3,43,4
(2,12)(2,1^{2}) [3,15][3,15] (3,22)(3,2^{2}) [3,36][3,36]
(3,12)(3,1^{2}) [6,36][6,36] (3,2,12)(3,2,1^{2}) [0,20][0,20]
(22,1)(2^{2},1) [3,20][3,20] (23,1)(2^{3},1) [0,4][0,4]
(2,13)(2,1^{3}) [0,4][0,4] (6,12)(6,1^{2}) [21,189][21,189]
(4,12)(4,1^{2}) [10,70][10,70] (5,2,1)(5,2,1) [24,256][24,256]
(3,2,1)(3,2,1) [8,64][8,64] (5,13)(5,1^{3}) [0,35][0,35]
(3,13)(3,1^{3}) [0,10][0,10] (4,3,1)(4,3,1) [15,175][15,175]
(23)(2^{3}) [1,10][1,10] (4,22)(4,2^{2}) [6,84][6,84]
(22,12)(2^{2},1^{2}) [0,6][0,6] (4,2,12)(4,2,1^{2}) [0,45][0,45]
(5,12)(5,1^{2}) [15,120][15,120] (32,2)(3^{2},2) [3,45][3,45]
(4,2,1)(4,2,1) [15,140][15,140] (32,12)(3^{2},1^{2}) [0,20][0,20]
(4,13)(4,1^{3}) [0,20][0,20] (3,22,1)(3,2^{2},1) [0,15][0,15]
(32,1)(3^{2},1) [6,60][6,60] (24)(2^{4}) [0,1][0,1]
Table 8: Ranks of Schur functors
𝔖10\mathfrak{S}_{10}-irreducible Missing Schur functors
(8,2)(8,2) 𝕊(3,12)+𝕊(2,14)\mathbb{S}_{(3,1^{2})}+\mathbb{S}_{(2,1^{4})}
(6,3,1)(6,3,1) 2𝕊(2,12)+𝕊(2,13)2\mathbb{S}_{(2,1^{2})}+\mathbb{S}_{(2,1^{3})}
(42,2)(4^{2},2) 0
(3,2,15)(3,2,1^{5}) 0
Table 9: Remaining corrections to the lower bounds

However, this is not always the case, e.g. for partition (8,2)(8,2) in our example. Let us explain why the remaining combination of Schurs occuring in our examples are as presented in Table 9. To gain more insight one can take the following steps:

  • As before, with 𝕊(2,1)\mathbb{S}_{(2,1)} having the smallest ratio between its first and third traces, now it is 𝕊(2,12)\mathbb{S}_{(2,1^{2})} that has the smallest ratio between its ranks in genus 33 and 44 (at least among the Schurs that are small enough to contribute nontrivially). Thus 𝕊(2,12)\mathbb{S}_{(2,1^{2})} must appear with multiplicity high enough to make the second column 6\leq 6 times the first. E.g. it appears at least once for partition (6,3,1)(6,3,1) leaving unknown ranks [3,19][3,19], at which point it is clear that it must appear with multiplicity 22 to account for the rank in genus 33. The remaining ranks [0,4][0,4] do not uniquely determine the multiplicities of Schur functors with 44 parts.

  • For partition (8,2)(8,2) ranks alone are not enough to uniquely determine the multiplicity of 𝕊(3,12)\mathbb{S}_{(3,1^{2})}. However, it is feasible to compute the trace of the matrix diag(2,1,1)diag(2,1,1), and this additional information is sufficient.

  • The second term of partition (8,2)(8,2), the functor 𝕊(2,14)\mathbb{S}_{(2,1^{4})}, does not even show up until one considers wedges of 5\geq 5 circles. However, since the representation χ(8,2)\chi_{(8,2)} has relatively small dimension, it is feasible to compute the nonequivariant rank of the isotypic component Hc9(F(X,10))(8,2)H^{9}_{c}(F(X,10))_{(8,2)} in genus 55 using the 22-term complex. This is sufficient for determining the 𝕊(3,12)\mathbb{S}_{(3,1^{2})} term from the previous point, and further detects that there exists some 𝕊λ\mathbb{S}_{\lambda} for λ\lambda with 55 parts (though it can not decide which).

With these additional considerations we have completely accounted for all Schur functors with 3\leq 3 parts and reduced the ambiguity regarding functors with 44 parts to a small number of cases of very small dimension: deciding between 𝕊(2,13),𝕊(23,1)\mathbb{S}_{(2,1^{3})},\mathbb{S}_{(2^{3},1)} and 𝕊(24)\mathbb{S}_{(2^{4})}.

Looking further to 𝕊λ\mathbb{S}_{\lambda} for λ\lambda with 5\geq 5 parts we reach the computational limits of the 2-term complex, where it is no longer feasible to compute further ranks or traces. Instead, observe that any yet undetected Schur functor must have relatively large polynomial degree. More precisely, any Schur functor with 5\geq 5 parts is either an exterior power or it has total polynomial degree 6\geq 6, as are the degrees of 𝕊(23,1)\mathbb{S}_{(2^{3},1)} and 𝕊(24)\mathbb{S}_{(2^{4})}.

Step 4: CE-complex in high polynomial degree

Fortunately, at high polynomial degree a new tool becomes available: high polynomial degrees appear as Φp1[10,10p]\Phi^{p-1}[10,10-p] for pp small, so the Chevalley–Eilenberg complex more readily gives access to the nonequivariant multiplicity of individual Schur functors. Although this complex forgets the data as to which 𝔖n\mathfrak{S}_{n}-isotypical component they belong, and further simplification is needed.

This couples with the crucial observation that by plugging wedges of even spheres into the CE-complex, the partitions of 𝕊λ\mathbb{S}_{\lambda} are the conjugate of the ones appearing for wedges of circles. For example the functor 𝕊(2,16)\mathbb{S}_{{(2,1^{6})}^{\ast}}, which would have been computationally expensive to access and require on a wedge of 7\geq 7 circles, contributes 𝕊(7,1)\mathbb{S}_{(7,1)} on a wedge of only two even spheres! And its high polynomial degree makes it even easier to access.

With this, the last step in our calculation is to compute the nonequivariant multiplicity of Schur functors of polynomial degree 6\geq 6 and compare them with the total multiplicity of those functors in our lower bound. If they are found to agree, the bound is sharp and all Schur functors have been counted. Fortunately, this always turned out to be the case

To make the last computations feasible we use one more trick: inputting wedges of spheres of different dimensions. Let us explain one such computation in detail, and for the other ones we will only specify what the analogous computation produces.

Step 4’: Mixing spheres of different dimensions

Say we want to compute the nonequivariant multiplicity of the Schur functor 𝕊(2,14)\mathbb{S}_{{(2,1^{4})}^{\ast}}, which on even spheres gives 𝕊(5,1)(H~(X))\mathbb{S}_{(5,1)}(\tilde{H}^{*}(X)) so two spheres suffice. Consider the space X=S2S4X=S^{2}\bigvee S^{4}.

The CE-complex for this XX admits a multigrading where the multidegree (a,b)(a,b) consists of tensors

(H2(S2)aH4(S4)b)(a+ba)H~(X)(a+b).(H^{2}(S^{2})^{\otimes a}\otimes H^{4}(S^{4})^{\otimes b})^{\oplus\binom{a+b}{a}}\leq\tilde{H}^{*}(X)^{\otimes(a+b)}.

This multigrading is preserved by the CE-differential, so one may decompose the complex and consider one multidegree at a time. The (a,b)(a,b)-multigraded subcomplex is substantially smaller than most 𝕊λ\mathbb{S}_{\lambda}-isotypic components, and furthermore it is spanned by a basis of pure tensors. These facts make calculations feasible, and we are able to determine the (a,b)(a,b)-multigraded part of the CE-homology.

At the same time, Maps(X,X)\operatorname{Maps}({X},{X}) acts on these subcomplexes, giving them the structure of polynomial GL(1)×GL(1)\operatorname{GL}(1)\times\operatorname{GL}(1)-representations, with the (a,b)(a,b)-multigraded component coinciding with the Syma(H2(X))Symb(H4(X))\operatorname{Sym}^{a}(H^{2}(X))\otimes\operatorname{Sym}^{b}(H^{4}(X))-isotypic component of this action. Knowing the multiplicity of each of these representations in homology, one is able to deduce the multiplicity of every 𝕊λ\mathbb{S}_{\lambda} for λ\lambda with 22 parts: one needs only to count how often every SymaSymb\operatorname{Sym}^{a}\otimes\operatorname{Sym}^{b} occurs in each 𝕊λ\mathbb{S}_{\lambda}.

Let us consider the (5,1)(5,1)-multigraded component. One can compute131313In general the multiplicity with which Syma1Symar\operatorname{Sym}^{a_{1}}\otimes\ldots\otimes\operatorname{Sym}^{a_{r}} appears in 𝕊λ\mathbb{S}_{\lambda} is equal to the number of semistandard Young tableax of shape λ\lambda filled with aia_{i} many ii’s. that Sym5Sym1\operatorname{Sym}^{5}\otimes\operatorname{Sym}^{1} only occurs once in 𝕊(5,1)\mathbb{S}_{(5,1)} and once in 𝕊(6)\mathbb{S}_{(6)}. It follows that its total multiplicity in CE-homology is Φ3[10,(5,1)]Φ3[10,(6)]\Phi^{3}[10,(5,1)]\oplus\Phi^{3}[10,(6)]. Given that we already know the multiplicities of all symmetric powers, this determines the rank of Φ3[10,(5,1)]\Phi^{3}[10,(5,1)].

One then compares this rank with the lower bound, and find that they agree. It follows that the bound accounts for the entire multiplicity of 𝕊(2,14)\mathbb{S}_{(2,1^{4})} on wedges of circles, which is thus known equivariantly. Note further that approaching this calculation with the multigraded components has the added benefit that it is sensitive to multiple Schur functors at a time: if the lower bounds of all of them sum to the calculated multiplicity of the multigraded component, then all their lower bounds in fact exhaust the whole multiplicity space.

Other cases of this approach needed to completely determine Hc9(F(X,10))H^{9}_{c}(F(X,10)) are:

  • X=S2S4X=S^{2}\bigvee S^{4}, multidegree (4,3)(4,3): computing rank of Φ2[10,(4,3)]+Φ2[10,(5,2)]+Φ2[10,(6,1)]+Φ2[10,(7)]\Phi^{2}[10,(4,3)]+\Phi^{2}[10,(5,2)]+\Phi^{2}[10,(6,1)]+\Phi^{2}[10,(7)].

  • X=S1S2X=S^{1}\bigvee S^{2}, multidegree (2,5)(2,5): computing rank of Φ2[10,(5,12)]+Φ2[10,(6,1)]\Phi^{2}[10,(5,1^{2})]+\Phi^{2}[10,(6,1)].

  • X=S2S4X=S^{2}\bigvee S^{4}, multidegree (4,4)(4,4): computing rank of Φ1[10,(4,4)]+Φ1[10,(5,3)]+Φ1[10,(6,2)]+Φ1[10,(7,1)]+Φ1[10,(8)]\Phi^{1}[10,(4,4)]+\Phi^{1}[10,(5,3)]+\Phi^{1}[10,(6,2)]+\Phi^{1}[10,(7,1)]+\Phi^{1}[10,(8)].

  • X=S2S4S6X=S^{2}\bigvee S^{4}\bigvee S^{6}, multidegree (5,2,1)(5,2,1): computing rank of Φ1[10,(5,2,1)]+Φ1[10,(5,3)]+Φ1[10,(6,1,1)]+2Φ1[10,(6,2)]+2Φ1[10,(7,1)]+Φ1[10,(8)]\Phi^{1}[10,(5,2,1)]+\Phi^{1}[10,(5,3)]+\Phi^{1}[10,(6,1,1)]+2\cdot\Phi^{1}[10,(6,2)]+2\cdot\Phi^{1}[10,(7,1)]+\Phi^{1}[10,(8)].

  • X=S1S2X=S^{1}\bigvee S^{2}, multidegree (3,5)(3,5): computing rank of Φ1[10,(5,13)]+Φ1[10,(6,12)]\Phi^{1}[10,(5,1^{3})]+\Phi^{1}[10,(6,1^{2})].

Note that in some of these cases we mix even and odd dimensional spheres. The partition (5,13)(5,1^{3}) has 44 parts, and its transpose has 55 parts. So it would take 4\geq 4 even spheres to access its multiplicity. But mixing even and odd spheres makes 𝕊(5,13)\mathbb{S}_{(5,1^{3})} contribute nontrivially already with only two spheres.

In all these cases the rank obtained from the CE-homology calculation agrees with the rank obtained from Table 10. Therefore all Schur functors 𝕊λ\mathbb{S}_{\lambda} for which Φ[10,λ]\Phi[10,{\lambda}^{\ast}] appeared in one of the cases above cannot occur with higher multiplicity in any 𝔖10\mathfrak{S}_{10}-isotypical component.

Careful bookkeeping shows that at this point every multiplicity of every Schur functor has been uniquely determined, therefore Table 10 contains the full decomposition of every isotypical component.

𝔖10\mathfrak{S}_{10}-irrep GL\operatorname{GL}-equivariant multiplicities in grHc9(F(X,10))\operatorname{gr}H^{9}_{c}(F(X,10))
(10)(10) 𝕊(4,1)+𝕊(3,13)+𝕊(2,15)+𝕊(18)\mathbb{S}_{(4,1)}+\mathbb{S}_{(3,1^{3})}+\mathbb{S}_{(2,1^{5})}+\mathbb{S}_{(1^{8})}
(9,1)(9,1) 𝕊(5)\mathbb{S}_{(5)}
(8,2)(8,2) 𝕊(4)+𝕊(3,12)+𝕊(3)+𝕊(2,14)+𝕊(2,1)+𝕊(2)+𝕊(17)+𝕊(14)+𝕊(13)\mathbb{S}_{(4)}+\mathbb{S}_{(3,1^{2})}+\mathbb{S}_{(3)}+\mathbb{S}_{(2,1^{4})}+\mathbb{S}_{(2,1)}+\mathbb{S}_{(2)}+\mathbb{S}_{(1^{7})}+\mathbb{S}_{(1^{4})}+\mathbb{S}_{(1^{3})}
(8,12)(8,1^{2}) 𝕊(4)+𝕊(3,1)+𝕊(3)+𝕊(2,13)+𝕊(2,12)+𝕊(2)+𝕊(16)+𝕊(15)+𝕊(12)+𝕊(1)\mathbb{S}_{(4)}+\mathbb{S}_{(3,1)}+\mathbb{S}_{(3)}+\mathbb{S}_{(2,1^{3})}+\mathbb{S}_{(2,1^{2})}+\mathbb{S}_{(2)}+\mathbb{S}_{(1^{6})}+\mathbb{S}_{(1^{5})}+\mathbb{S}_{(1^{2})}+\mathbb{S}_{(1)}
(7,3)(7,3) 𝕊(5)+4𝕊(3)+𝕊(2,12)+2𝕊(2,1)+2𝕊(2)+𝕊(15)+𝕊(14)+𝕊(13)+2𝕊(12)+𝕊(1)\mathbb{S}_{(5)}+4\mathbb{S}_{(3)}+\mathbb{S}_{(2,1^{2})}+2\mathbb{S}_{(2,1)}+2\mathbb{S}_{(2)}+\mathbb{S}_{(1^{5})}+\mathbb{S}_{(1^{4})}+\mathbb{S}_{(1^{3})}+2\mathbb{S}_{(1^{2})}+\mathbb{S}_{(1)}
(7,2,1)(7,2,1) 𝕊(5)+3𝕊(4)+𝕊(3,1)+5𝕊(3)+𝕊(2,13)+𝕊(2,12)+4𝕊(2,1)+5𝕊(2)+𝕊(16)+𝕊(15)+2𝕊(14)+3𝕊(13)+3𝕊(12)+2𝕊(1)\mathbb{S}_{(5)}+3\mathbb{S}_{(4)}+\mathbb{S}_{(3,1)}+5\mathbb{S}_{(3)}+\mathbb{S}_{(2,1^{3})}+\mathbb{S}_{(2,1^{2})}+4\mathbb{S}_{(2,1)}+5\mathbb{S}_{(2)}+\mathbb{S}_{(1^{6})}+\mathbb{S}_{(1^{5})}+2\mathbb{S}_{(1^{4})}+3\mathbb{S}_{(1^{3})}+3\mathbb{S}_{(1^{2})}+2\mathbb{S}_{(1)}
(7,13)(7,1^{3}) 𝕊(6)+2𝕊(4)+2𝕊(3)+𝕊(2,12)+𝕊(2,1)+3𝕊(2)+𝕊(15)+𝕊(14)+𝕊(13)+2𝕊(12)\mathbb{S}_{(6)}+2\mathbb{S}_{(4)}+2\mathbb{S}_{(3)}+\mathbb{S}_{(2,1^{2})}+\mathbb{S}_{(2,1)}+3\mathbb{S}_{(2)}+\mathbb{S}_{(1^{5})}+\mathbb{S}_{(1^{4})}+\mathbb{S}_{(1^{3})}+2\mathbb{S}_{(1^{2})}
(6,4)(6,4) 2𝕊(4)+2𝕊(3)+2𝕊(2,1)+4𝕊(2)+𝕊(14)+3𝕊(13)+2𝕊(12)2\mathbb{S}_{(4)}+2\mathbb{S}_{(3)}+2\mathbb{S}_{(2,1)}+4\mathbb{S}_{(2)}+\mathbb{S}_{(1^{4})}+3\mathbb{S}_{(1^{3})}+2\mathbb{S}_{(1^{2})}
(6,3,1)(6,3,1) 𝕊(5)+6𝕊(4)+2𝕊(3,1)+11𝕊(3)+𝕊(2,13)+2𝕊(2,12)+7𝕊(2,1)+10𝕊(2)+𝕊(16)+2𝕊(15)+3𝕊(14)+5𝕊(13)+7𝕊(12)+4𝕊(1)\mathbb{S}_{(5)}+6\mathbb{S}_{(4)}+2\mathbb{S}_{(3,1)}+11\mathbb{S}_{(3)}+\mathbb{S}_{(2,1^{3})}+2\mathbb{S}_{(2,1^{2})}+7\mathbb{S}_{(2,1)}+10\mathbb{S}_{(2)}+\mathbb{S}_{(1^{6})}+2\mathbb{S}_{(1^{5})}+3\mathbb{S}_{(1^{4})}+5\mathbb{S}_{(1^{3})}+7\mathbb{S}_{(1^{2})}+4\mathbb{S}_{(1)}
(6,22)(6,2^{2}) 𝕊(5)+4𝕊(4)+7𝕊(3)+6𝕊(2,1)+8𝕊(2)+3𝕊(14)+6𝕊(13)+4𝕊(12)+𝕊(1)\mathbb{S}_{(5)}+4\mathbb{S}_{(4)}+7\mathbb{S}_{(3)}+6\mathbb{S}_{(2,1)}+8\mathbb{S}_{(2)}+3\mathbb{S}_{(1^{4})}+6\mathbb{S}_{(1^{3})}+4\mathbb{S}_{(1^{2})}+\mathbb{S}_{(1)}
(6,2,12)(6,2,1^{2}) 2𝕊(5)+6𝕊(4)+𝕊(3,1)+13𝕊(3)+2𝕊(2,12)+8𝕊(2,1)+10𝕊(2)+2𝕊(15)+3𝕊(14)+5𝕊(13)+7𝕊(12)+5𝕊(1)2\mathbb{S}_{(5)}+6\mathbb{S}_{(4)}+\mathbb{S}_{(3,1)}+13\mathbb{S}_{(3)}+2\mathbb{S}_{(2,1^{2})}+8\mathbb{S}_{(2,1)}+10\mathbb{S}_{(2)}+2\mathbb{S}_{(1^{5})}+3\mathbb{S}_{(1^{4})}+5\mathbb{S}_{(1^{3})}+7\mathbb{S}_{(1^{2})}+5\mathbb{S}_{(1)}
(6,14)(6,1^{4}) 𝕊(5)+2𝕊(4)+4𝕊(3)+3𝕊(2,1)+4𝕊(2)+2𝕊(14)+3𝕊(13)+2𝕊(12)+𝕊(1)\mathbb{S}_{(5)}+2\mathbb{S}_{(4)}+4\mathbb{S}_{(3)}+3\mathbb{S}_{(2,1)}+4\mathbb{S}_{(2)}+2\mathbb{S}_{(1^{4})}+3\mathbb{S}_{(1^{3})}+2\mathbb{S}_{(1^{2})}+\mathbb{S}_{(1)}
(52)(5^{2}) 𝕊(5)+3𝕊(3)+3𝕊(2,1)+𝕊(14)+𝕊(13)+2𝕊(1)\mathbb{S}_{(5)}+3\mathbb{S}_{(3)}+3\mathbb{S}_{(2,1)}+\mathbb{S}_{(1^{4})}+\mathbb{S}_{(1^{3})}+2\mathbb{S}_{(1)}
(5,4,1)(5,4,1) 𝕊(5)+5𝕊(4)+𝕊(3,1)+10𝕊(3)+𝕊(2,12)+7𝕊(2,1)+9𝕊(2)+𝕊(15)+3𝕊(14)+5𝕊(13)+6𝕊(12)+3𝕊(1)\mathbb{S}_{(5)}+5\mathbb{S}_{(4)}+\mathbb{S}_{(3,1)}+10\mathbb{S}_{(3)}+\mathbb{S}_{(2,1^{2})}+7\mathbb{S}_{(2,1)}+9\mathbb{S}_{(2)}+\mathbb{S}_{(1^{5})}+3\mathbb{S}_{(1^{4})}+5\mathbb{S}_{(1^{3})}+6\mathbb{S}_{(1^{2})}+3\mathbb{S}_{(1)}
(5,3,2)(5,3,2) 2𝕊(5)+7𝕊(4)+𝕊(3,1)+18𝕊(3)+𝕊(2,12)+12𝕊(2,1)+13𝕊(2)+𝕊(15)+4𝕊(14)+7𝕊(13)+9𝕊(12)+6𝕊(1)2\mathbb{S}_{(5)}+7\mathbb{S}_{(4)}+\mathbb{S}_{(3,1)}+18\mathbb{S}_{(3)}+\mathbb{S}_{(2,1^{2})}+12\mathbb{S}_{(2,1)}+13\mathbb{S}_{(2)}+\mathbb{S}_{(1^{5})}+4\mathbb{S}_{(1^{4})}+7\mathbb{S}_{(1^{3})}+9\mathbb{S}_{(1^{2})}+6\mathbb{S}_{(1)}
(5,3,12)(5,3,1^{2}) 𝕊(6)+𝕊(5)+13𝕊(4)+2𝕊(3,1)+16𝕊(3)+𝕊(2,12)+10𝕊(2,1)+20𝕊(2)+𝕊(15)+4𝕊(14)+10𝕊(13)+12𝕊(12)+5𝕊(1)\mathbb{S}_{(6)}+\mathbb{S}_{(5)}+13\mathbb{S}_{(4)}+2\mathbb{S}_{(3,1)}+16\mathbb{S}_{(3)}+\mathbb{S}_{(2,1^{2})}+10\mathbb{S}_{(2,1)}+20\mathbb{S}_{(2)}+\mathbb{S}_{(1^{5})}+4\mathbb{S}_{(1^{4})}+10\mathbb{S}_{(1^{3})}+12\mathbb{S}_{(1^{2})}+5\mathbb{S}_{(1)}
(5,22,1)(5,2^{2},1) 4𝕊(5)+8𝕊(4)+𝕊(3,1)+19𝕊(3)+𝕊(2,12)+11𝕊(2,1)+16𝕊(2)+𝕊(15)+3𝕊(14)+8𝕊(13)+11𝕊(12)+6𝕊(1)4\mathbb{S}_{(5)}+8\mathbb{S}_{(4)}+\mathbb{S}_{(3,1)}+19\mathbb{S}_{(3)}+\mathbb{S}_{(2,1^{2})}+11\mathbb{S}_{(2,1)}+16\mathbb{S}_{(2)}+\mathbb{S}_{(1^{5})}+3\mathbb{S}_{(1^{4})}+8\mathbb{S}_{(1^{3})}+11\mathbb{S}_{(1^{2})}+6\mathbb{S}_{(1)}
(5,2,13)(5,2,1^{3}) 𝕊(6)+3𝕊(5)+8𝕊(4)+14𝕊(3)+9𝕊(2,1)+13𝕊(2)+3𝕊(14)+6𝕊(13)+8𝕊(12)+5𝕊(1)\mathbb{S}_{(6)}+3\mathbb{S}_{(5)}+8\mathbb{S}_{(4)}+14\mathbb{S}_{(3)}+9\mathbb{S}_{(2,1)}+13\mathbb{S}_{(2)}+3\mathbb{S}_{(1^{4})}+6\mathbb{S}_{(1^{3})}+8\mathbb{S}_{(1^{2})}+5\mathbb{S}_{(1)}
(5,15)(5,1^{5}) 𝕊(7)+2𝕊(5)+𝕊(4)+5𝕊(3)+2𝕊(2,1)+3𝕊(2)+𝕊(13)+2𝕊(12)+2𝕊(1)\mathbb{S}_{(7)}+2\mathbb{S}_{(5)}+\mathbb{S}_{(4)}+5\mathbb{S}_{(3)}+2\mathbb{S}_{(2,1)}+3\mathbb{S}_{(2)}+\mathbb{S}_{(1^{3})}+2\mathbb{S}_{(1^{2})}+2\mathbb{S}_{(1)}
(42,2)(4^{2},2) 𝕊(5)+6𝕊(4)+𝕊(3,1)+6𝕊(3)+𝕊(22)+4𝕊(2,1)+10𝕊(2)+2𝕊(14)+6𝕊(13)+5𝕊(12)+𝕊(1)\mathbb{S}_{(5)}+6\mathbb{S}_{(4)}+\mathbb{S}_{(3,1)}+6\mathbb{S}_{(3)}+\mathbb{S}_{(2^{2})}+4\mathbb{S}_{(2,1)}+10\mathbb{S}_{(2)}+2\mathbb{S}_{(1^{4})}+6\mathbb{S}_{(1^{3})}+5\mathbb{S}_{(1^{2})}+\mathbb{S}_{(1)}
(42,12)(4^{2},1^{2}) 3𝕊(5)+3𝕊(4)+14𝕊(3)+𝕊(2,12)+8𝕊(2,1)+7𝕊(2)+𝕊(15)+2𝕊(14)+3𝕊(13)+6𝕊(12)+5𝕊(1)3\mathbb{S}_{(5)}+3\mathbb{S}_{(4)}+14\mathbb{S}_{(3)}+\mathbb{S}_{(2,1^{2})}+8\mathbb{S}_{(2,1)}+7\mathbb{S}_{(2)}+\mathbb{S}_{(1^{5})}+2\mathbb{S}_{(1^{4})}+3\mathbb{S}_{(1^{3})}+6\mathbb{S}_{(1^{2})}+5\mathbb{S}_{(1)}
(4,32)(4,3^{2}) 𝕊(5)+4𝕊(4)+𝕊(3,1)+8𝕊(3)+𝕊(2,12)+4𝕊(2,1)+6𝕊(2)+𝕊(15)+𝕊(14)+2𝕊(13)+5𝕊(12)+3𝕊(1)\mathbb{S}_{(5)}+4\mathbb{S}_{(4)}+\mathbb{S}_{(3,1)}+8\mathbb{S}_{(3)}+\mathbb{S}_{(2,1^{2})}+4\mathbb{S}_{(2,1)}+6\mathbb{S}_{(2)}+\mathbb{S}_{(1^{5})}+\mathbb{S}_{(1^{4})}+2\mathbb{S}_{(1^{3})}+5\mathbb{S}_{(1^{2})}+3\mathbb{S}_{(1)}
(4,3,2,1)(4,3,2,1) 4𝕊(5)+15𝕊(4)+2𝕊(3,1)+25𝕊(3)+16𝕊(2,1)+23𝕊(2)+4𝕊(14)+11𝕊(13)+14𝕊(12)+9𝕊(1)4\mathbb{S}_{(5)}+15\mathbb{S}_{(4)}+2\mathbb{S}_{(3,1)}+25\mathbb{S}_{(3)}+16\mathbb{S}_{(2,1)}+23\mathbb{S}_{(2)}+4\mathbb{S}_{(1^{4})}+11\mathbb{S}_{(1^{3})}+14\mathbb{S}_{(1^{2})}+9\mathbb{S}_{(1)}
(4,3,13)(4,3,1^{3}) 𝕊(6)+4𝕊(5)+9𝕊(4)+17𝕊(3)+9𝕊(2,1)+16𝕊(2)+2𝕊(14)+7𝕊(13)+10𝕊(12)+5𝕊(1)\mathbb{S}_{(6)}+4\mathbb{S}_{(5)}+9\mathbb{S}_{(4)}+17\mathbb{S}_{(3)}+9\mathbb{S}_{(2,1)}+16\mathbb{S}_{(2)}+2\mathbb{S}_{(1^{4})}+7\mathbb{S}_{(1^{3})}+10\mathbb{S}_{(1^{2})}+5\mathbb{S}_{(1)}
(4,23)(4,2^{3}) 𝕊(5)+7𝕊(4)+𝕊(3,1)+8𝕊(3)+4𝕊(2,1)+11𝕊(2)+𝕊(14)+5𝕊(13)+6𝕊(12)+𝕊(1)\mathbb{S}_{(5)}+7\mathbb{S}_{(4)}+\mathbb{S}_{(3,1)}+8\mathbb{S}_{(3)}+4\mathbb{S}_{(2,1)}+11\mathbb{S}_{(2)}+\mathbb{S}_{(1^{4})}+5\mathbb{S}_{(1^{3})}+6\mathbb{S}_{(1^{2})}+\mathbb{S}_{(1)}
(4,22,12)(4,2^{2},1^{2}) 𝕊(6)+4𝕊(5)+11𝕊(4)+𝕊(3,1)+19𝕊(3)+9𝕊(2,1)+16𝕊(2)+𝕊(14)+5𝕊(13)+11𝕊(12)+8𝕊(1)\mathbb{S}_{(6)}+4\mathbb{S}_{(5)}+11\mathbb{S}_{(4)}+\mathbb{S}_{(3,1)}+19\mathbb{S}_{(3)}+9\mathbb{S}_{(2,1)}+16\mathbb{S}_{(2)}+\mathbb{S}_{(1^{4})}+5\mathbb{S}_{(1^{3})}+11\mathbb{S}_{(1^{2})}+8\mathbb{S}_{(1)}
(4,2,14)(4,2,1^{4}) 2𝕊(6)+3𝕊(5)+7𝕊(4)+8𝕊(3)+3𝕊(2,1)+11𝕊(2)+4𝕊(13)+6𝕊(12)+3𝕊(1)2\mathbb{S}_{(6)}+3\mathbb{S}_{(5)}+7\mathbb{S}_{(4)}+8\mathbb{S}_{(3)}+3\mathbb{S}_{(2,1)}+11\mathbb{S}_{(2)}+4\mathbb{S}_{(1^{3})}+6\mathbb{S}_{(1^{2})}+3\mathbb{S}_{(1)}
(4,16)(4,1^{6}) 𝕊(6)+𝕊(5)+𝕊(4)+2𝕊(3)+2𝕊(2)+2𝕊(12)+𝕊(1)\mathbb{S}_{(6)}+\mathbb{S}_{(5)}+\mathbb{S}_{(4)}+2\mathbb{S}_{(3)}+2\mathbb{S}_{(2)}+2\mathbb{S}_{(1^{2})}+\mathbb{S}_{(1)}
(33,1)(3^{3},1) 𝕊(6)+5𝕊(4)+5𝕊(3)+2𝕊(2,1)+8𝕊(2)+3𝕊(13)+5𝕊(12)+𝕊(1)\mathbb{S}_{(6)}+5\mathbb{S}_{(4)}+5\mathbb{S}_{(3)}+2\mathbb{S}_{(2,1)}+8\mathbb{S}_{(2)}+3\mathbb{S}_{(1^{3})}+5\mathbb{S}_{(1^{2})}+\mathbb{S}_{(1)}
(32,22)(3^{2},2^{2}) 3𝕊(5)+3𝕊(4)+12𝕊(3)+7𝕊(2,1)+5𝕊(2)+𝕊(14)+2𝕊(13)+4𝕊(12)+5𝕊(1)3\mathbb{S}_{(5)}+3\mathbb{S}_{(4)}+12\mathbb{S}_{(3)}+7\mathbb{S}_{(2,1)}+5\mathbb{S}_{(2)}+\mathbb{S}_{(1^{4})}+2\mathbb{S}_{(1^{3})}+4\mathbb{S}_{(1^{2})}+5\mathbb{S}_{(1)}
(32,2,12)(3^{2},2,1^{2}) 𝕊(6)+3𝕊(5)+9𝕊(4)+𝕊(3,1)+13𝕊(3)+6𝕊(2,1)+14𝕊(2)+𝕊(14)+5𝕊(13)+8𝕊(12)+4𝕊(1)\mathbb{S}_{(6)}+3\mathbb{S}_{(5)}+9\mathbb{S}_{(4)}+\mathbb{S}_{(3,1)}+13\mathbb{S}_{(3)}+6\mathbb{S}_{(2,1)}+14\mathbb{S}_{(2)}+\mathbb{S}_{(1^{4})}+5\mathbb{S}_{(1^{3})}+8\mathbb{S}_{(1^{2})}+4\mathbb{S}_{(1)}
(32,14)(3^{2},1^{4}) 𝕊(7)+4𝕊(5)+2𝕊(4)+9𝕊(3)+4𝕊(2,1)+4𝕊(2)+𝕊(13)+3𝕊(12)+4𝕊(1)\mathbb{S}_{(7)}+4\mathbb{S}_{(5)}+2\mathbb{S}_{(4)}+9\mathbb{S}_{(3)}+4\mathbb{S}_{(2,1)}+4\mathbb{S}_{(2)}+\mathbb{S}_{(1^{3})}+3\mathbb{S}_{(1^{2})}+4\mathbb{S}_{(1)}
(3,23,1)(3,2^{3},1) 3𝕊(5)+6𝕊(4)+𝕊(3,1)+8𝕊(3)+3𝕊(2,1)+8𝕊(2)+2𝕊(13)+5𝕊(12)+3𝕊(1)3\mathbb{S}_{(5)}+6\mathbb{S}_{(4)}+\mathbb{S}_{(3,1)}+8\mathbb{S}_{(3)}+3\mathbb{S}_{(2,1)}+8\mathbb{S}_{(2)}+2\mathbb{S}_{(1^{3})}+5\mathbb{S}_{(1^{2})}+3\mathbb{S}_{(1)}
(3,22,13)(3,2^{2},1^{3}) 2𝕊(6)+2𝕊(5)+7𝕊(4)+8𝕊(3)+3𝕊(2,1)+9𝕊(2)+2𝕊(13)+5𝕊(12)+3𝕊(1)2\mathbb{S}_{(6)}+2\mathbb{S}_{(5)}+7\mathbb{S}_{(4)}+8\mathbb{S}_{(3)}+3\mathbb{S}_{(2,1)}+9\mathbb{S}_{(2)}+2\mathbb{S}_{(1^{3})}+5\mathbb{S}_{(1^{2})}+3\mathbb{S}_{(1)}
(3,2,15)(3,2,1^{5}) 𝕊(7)+𝕊(6)+2𝕊(5)+3𝕊(4)+3𝕊(3)+4𝕊(2)+2𝕊(12)+2𝕊(1)\mathbb{S}_{(7)}+\mathbb{S}_{(6)}+2\mathbb{S}_{(5)}+3\mathbb{S}_{(4)}+3\mathbb{S}_{(3)}+4\mathbb{S}_{(2)}+2\mathbb{S}_{(1^{2})}+2\mathbb{S}_{(1)}
(3,17)(3,1^{7}) 𝕊(8)+𝕊(6)+𝕊(4)+𝕊(2)\mathbb{S}_{(8)}+\mathbb{S}_{(6)}+\mathbb{S}_{(4)}+\mathbb{S}_{(2)}
(25)(2^{5}) 2𝕊(4)+2𝕊(2)+𝕊(13)2\mathbb{S}_{(4)}+2\mathbb{S}_{(2)}+\mathbb{S}_{(1^{3})}
(24,12)(2^{4},1^{2}) 2𝕊(5)+𝕊(4)+4𝕊(3)+2𝕊(2,1)+𝕊(2)+𝕊(12)+2𝕊(1)2\mathbb{S}_{(5)}+\mathbb{S}_{(4)}+4\mathbb{S}_{(3)}+2\mathbb{S}_{(2,1)}+\mathbb{S}_{(2)}+\mathbb{S}_{(1^{2})}+2\mathbb{S}_{(1)}
(23,14)(2^{3},1^{4}) 𝕊(6)+𝕊(5)+𝕊(4)+2𝕊(3)+2𝕊(2)+𝕊(12)\mathbb{S}_{(6)}+\mathbb{S}_{(5)}+\mathbb{S}_{(4)}+2\mathbb{S}_{(3)}+2\mathbb{S}_{(2)}+\mathbb{S}_{(1^{2})}
(22,16)(2^{2},1^{6}) 𝕊(7)+𝕊(5)+𝕊(3)+𝕊(1)\mathbb{S}_{(7)}+\mathbb{S}_{(5)}+\mathbb{S}_{(3)}+\mathbb{S}_{(1)}
(2,18)(2,1^{8}) 0
(110)(1^{10}) 𝕊(9)\mathbb{S}_{(9)}
Table 10: Equivariant multiplicity of each isotypic component in grHc9(F(X,10))\operatorname{gr}H_{c}^{9}(F(X,10))

Appendix B Conjectural generating functions

We present here the conjectural multiplicities from Conjectures 6.9 and 6.13 in the form of generating functions.

Let Λ\Lambda denote the ring of symmetric functions (see e.g. [Mac95]). Some of the generating functions below are expressed as elements of Λ[[t]]\Lambda[[t]], the λ\lambda-ring of power series in tt with coefficients in Λ\Lambda. Let ch:K0(FunPoly)Λ\operatorname{ch}:K_{0}(\operatorname{Fun_{Poly}})\to\Lambda be the linear function on the Grothendieck group of polynomial functors, defined by sending the Schur functor 𝕊λ\mathbb{S}_{\lambda} to the Schur symmetric function sλs_{\lambda}, and let Exp:tΛ[[t]]Λ[[t]]\operatorname{Exp}\colon t\Lambda[[t]]\to\Lambda[[t]] be the plethystic exponential, defined e.g. in [GK98].

Consider partitions with one long row λ[n]=(n|λ|,λ)\lambda[n]=(n-|\lambda|,\lambda). The generating functions

n1(t)nchGL(g)(grHcn1(F(gS1,n))𝔖nχλ[n])and\displaystyle\sum_{n\geq 1}{(-t)^{n}\cdot\operatorname{ch}_{\operatorname{GL}(g)}\left(\operatorname{gr}H_{c}^{n-1}(F(\bigvee_{g}S^{1},n))\otimes_{\mathfrak{S}_{n}}\chi_{\lambda[n]}\right)}\quad\text{and}
n1tnχλ[n],gr0WHcn+2(2,n,)𝔖n\displaystyle\sum_{n\geq 1}{t^{n}\cdot\langle\chi_{\lambda[n]},\ \operatorname{gr}_{0}^{W}H_{c}^{n+2}(\mathcal{M}_{2,n},\mathbb{Q})\rangle_{\mathfrak{S}_{n}}}

are, respectively,

  • for λ[n]=(n)\lambda[n]=(n) they are t2(Exp(s1(t2t))1)1tt\frac{t^{2}(\operatorname{Exp}(s_{1}(t^{2}-t))-1)}{1-t}-t and (t4t7+t10)(1+t3)(1t4)(1t12)\frac{(t^{4}-t^{7}+t^{10})(1+t^{3})}{(1-t^{4})(1-t^{12})} (see 6.8),

  • for λ[n]=(n1,1)\lambda[n]=(n-1,1) they are Exp(s1t2)1\operatorname{Exp}(s_{1}t^{2})-1 and t12(1t4)(1t12)\frac{t^{12}}{(1-t^{4})(1-t^{12})} (see 6.12),

  • for λ[n]=(n2,2)\lambda[n]=(n-2,2) they are 11t4((tt4)(1+s1t2Exp(s1t2))(t3+t4+t5)(Exp(s1(t2t))1))\frac{1}{1-t^{4}}\Big{(}(t-t^{4})(1+s_{1}t^{2}-\operatorname{Exp}(s_{1}t^{2}))-(t^{3}+t^{4}+t^{5})\cdot({\operatorname{Exp}(s_{1}(t^{2}-t))}-1)\Big{)} and (t5+t13+t14)(1+t3)(1t4)2(1t12)\frac{(t^{5}+t^{13}+t^{14})(1+t^{3})}{(1-t^{4})^{2}(1-t^{12})},

  • for λ[n]=(n2,1,1)\lambda[n]=(n-2,1,1) they are 11t4((1t3)1+s1t2Exp(s1t2)t+t4(Exp(s1(t2t))1))\frac{1}{1-t^{4}}\left((1-t^{3})\frac{1+s_{1}t^{2}-\operatorname{Exp}(s_{1}t^{2})}{t}+t^{4}(\operatorname{Exp}(s_{1}(t^{2}-t))-1)\right) and (t6t8+t11)(1t6)(1t)(1t4)2(1t12)\frac{(t^{6}-t^{8}+t^{11})(1-t^{6})}{(1-t)(1-t^{4})^{2}(1-t^{12})}.

For partitions with one long column λ[n]{\lambda[n]}^{\ast} the generating functions

n1tnchGL(g)(grHcn1(F(gS1,n))𝔖nχλ[n])and\displaystyle\sum_{n\geq 1}{t^{n}\cdot\operatorname{ch}_{\operatorname{GL}(g)}\left(\operatorname{gr}H_{c}^{n-1}(F(\bigvee_{g}S^{1},n))\otimes_{\mathfrak{S}_{n}}\chi_{{\lambda[n]}^{\ast}}\right)}\quad\text{and}
n1tnχλ[n],gr0WHcn+2(2,n,)𝔖n\displaystyle\sum_{n\geq 1}{t^{n}\cdot\langle\chi_{{\lambda[n]}^{\ast}},\ \operatorname{gr}_{0}^{W}H_{c}^{n+2}(\mathcal{M}_{2,n},\mathbb{Q})\rangle_{\mathfrak{S}_{n}}}

are, respectively,

  • for λ[n]=(1n){\lambda[n]}^{\ast}=(1^{n}) they are tExp(s1t)t\operatorname{Exp}(s_{1}t) and t7(1t2)(1t6)\frac{t^{7}}{(1-t^{2})(1-t^{6})} (see 6.8)

  • for λ[n]=(2,1n1){\lambda[n]}^{\ast}=(2,1^{n-1}) they are 0 for both generating functions (see 6.8),

  • for λ[n]=(2,2,1n4){\lambda[n]}^{\ast}=(2,2,1^{n-4}) they are t3(Exp(s1t)1)1t2\frac{t^{3}(\operatorname{Exp}(s_{1}t)-1)}{1-t^{2}} and t9(1t2)2(1t6)\frac{t^{9}}{(1-t^{2})^{2}(1-t^{6})}

  • for λ[n]=(3,1n3){\lambda[n]}^{\ast}=(3,1^{n-3}) they are t2(Exp(s1t)1)1t2\frac{t^{2}(\operatorname{Exp}(s_{1}t)-1)}{1-t^{2}} and t8(1t2)2(1t6)\frac{t^{8}}{(1-t^{2})^{2}(1-t^{6})}

  • for λ[n]=(2,2,2,1n6){\lambda[n]}^{\ast}=(2,2,2,1^{n-6}) they are t4(Exp(s1t)1s1t)(1t)(1t3)+t7(s1t2+s1,1)(1t2)(1t3)\frac{t^{4}(\operatorname{Exp}(s_{1}t)-1-s_{1}t)}{(1-t)(1-t^{3})}+\frac{t^{7}(s_{1}t^{2}+s_{1,1})}{(1-t^{2})(1-t^{3})} and t7t8+t10t13+t14(1t)(1t2)(1t3)(1t6)\frac{t^{7}-t^{8}+t^{10}-t^{13}+t^{14}}{(1-t)(1-t^{2})(1-t^{3})(1-t^{6})}

  • for λ[n]=(3,2,1n5){\lambda[n]}^{\ast}=(3,2,1^{n-5}) they are (t3+t5)(Exp(s1t)1s1t)+t5(s1t2+s1,1)(1t)(1t3)\frac{(t^{3}+t^{5})(\operatorname{Exp}(s_{1}t)-1-s_{1}t)+t^{5}(s_{1}t^{2}+s_{1,1})}{(1-t)(1-t^{3})} and t5t7+t9+t13(1t)(1t2)(1t3)(1t6)\frac{t^{5}-t^{7}+t^{9}+t^{13}}{(1-t)(1-t^{2})(1-t^{3})(1-t^{6})}

  • for λ[n]=(4,1n4){\lambda[n]}^{\ast}=(4,1^{n-4}) they are t4(Exp(s1t)1s1t)(1t)(1t3)+t4(s1t2+s1,1)(1t2)(1t3)\frac{t^{4}(\operatorname{Exp}(s_{1}t)-1-s_{1}t)}{(1-t)(1-t^{3})}+\frac{t^{4}(s_{1}t^{2}+s_{1,1})}{(1-t^{2})(1-t^{3})} and t4t5+t11(1t)(1t2)(1t3)(1t6)\frac{t^{4}-t^{5}+t^{11}}{(1-t)(1-t^{2})(1-t^{3})(1-t^{6})}.

Remark B.1.

The agreement of these expressions with the conjectural characters in Conjecture 6.9 has been verified computationally up to a high degree of confidence. However, we have not proved this formally.

References

  • [Ada97] Victor Adamchik “On Stirling numbers and Euler sums” In Journal of Computational and Applied Mathematics 79.1, 1997, pp. 119–130 DOI: https://doi.org/10.1016/S0377-0427(96)00167-7
  • [AT14] Gregory Arone and Victor Turchin “On the rational homology of high-dimensional analogues of spaces of long knots” In Geometry & Topology 18.3 Mathematical Sciences Publishers, 2014, pp. 1261–1322 DOI: 10.2140/gt.2014.18.1261
  • [BCGY21] Christin Bibby, Melody Chan, Nir Gadish and Claudia He Yun “Homology representations of compactified configurations on graphs applied to M2,n{M_{2,n}}”, 2021 arXiv:2109.03302v4
  • [BG19] Christin Bibby and Nir Gadish “A generating function approach to new representation stability phenomena in orbit configuration spaces”, 2019 arXiv:1911.02125v2
  • [BBP20] Tara Brendle, Nathan Broaddus and Andrew Putman “The high-dimensional cohomology of the moduli space of curves with level structures II: punctures and boundary” to appear in Israel J. Math, 2020 arXiv:2003.10913v5
  • [CFGP19] Melody Chan, Carel Faber, Søren Galatius and Sam Payne “The Sn{S}_{n}-equivariant top weight Euler characteristic of Mg,n{M}_{g,n}”, 2019 arXiv:1904.06367v2
  • [CGP22] Melody Chan, Søren Galatius and Sam Payne “Topology of Moduli Spaces of Tropical Curves with Marked Points” In Facets of Algebraic Geometry: A Collection in Honor of William Fulton’s 80th Birthday 1, London Mathematical Society Lecture Note Series Cambridge University Press, 2022, pp. 77–131 DOI: 10.1017/9781108877831.004
  • [Coh76] Fred Cohen “The homology of Cn+1C_{n+1}-Spaces, n0n\geq 0 In The Homology of Iterated Loop Spaces Berlin, Heidelberg: Springer Berlin Heidelberg, 1976, pp. 207–351 DOI: 10.1007/BFb0080467
  • [CT78] Frederick R Cohen and Laurence R Taylor “Computations of Gelfand-Fuks cohomology, the cohomology of function spaces, and the cohomology of configuration spaces” In Geometric applications of homotopy theory I Springer, 1978, pp. 106–143
  • [DV15] Aurélien Djament and Christine Vespa “Sur l’homologie des groupes d’automorphismes des groupes libres à coefficients polynomiaux” In Commentarii Mathematici Helvetici 90.1, 2015, pp. 33–58 DOI: 10.4171/CMH/345
  • [ER19] Nicholas Early and Victor Reiner “On configuration spaces and Whitehouse’s lifts of the Eulerian representations” In Journal of Pure and Applied Algebra 223.10 Elsevier, 2019, pp. 4524–4535 DOI: 10.1016/j.jpaa.2019.02.002
  • [EM54] Samuel Eilenberg and Saunders MacLane “On the Groups H(Π,n)H(\Pi,n), II: Methods of Computation” In Annals of Mathematics 60.1 Annals of Mathematics, 1954, pp. 49–139 DOI: 10.2307/1969702
  • [Erd53] Paul Erdös “On a conjecture of Hammersley” In Journal of the London Mathematical Society 1.2 Wiley Online Library, 1953, pp. 232–236 DOI: 10.1112/jlms/s1-28.2.232
  • [FZ00] Eva Maria Feichtner and Günter M Ziegler “The integral cohomology algebras of ordered configuration spaces of spheres” In Documenta Mathematica 5, 2000, pp. 115–139
  • [Get95] Ezra Getzler “Operads and Moduli Spaces of Genus 0 Riemann Surfaces” In The Moduli Space of Curves Boston, MA: Birkhäuser Boston, 1995, pp. 199–230 DOI: 10.1007/978-1-4612-4264-2_8
  • [Get99] Ezra Getzler “Resolving mixed Hodge modules on configuration spaces” In Duke Mathematical Journal 96.1 Duke University Press, 1999, pp. 175–203 DOI: 10.1215/S0012-7094-99-09605-9
  • [Get99a] Ezra Getzler “The homology groups of some two-step nilpotent Lie algebras associated to symplectic vector spaces”, 1999 arXiv:math/9903147v1
  • [GK98] Ezra Getzler and Mikhail M. Kapranov “Modular operads” In Compositio Mathematica 110.1 London Mathematical Society, 1998, pp. 65–125 DOI: 10.1023/A:1000245600345
  • [GK02] Robert W Ghrist and Daniel E Koditschek “Safe cooperative robot dynamics on graphs” In SIAM journal on control and optimization 40.5 SIAM, 2002, pp. 1556–1575 DOI: 10.1137/S0363012900368442
  • [HZ86] John Harer and Don Zagier “The Euler characteristic of the moduli space of curves.” In Inventiones mathematicae 85 Springer-Verlag, Berlin, 1986, pp. 457–485 DOI: 10.1007/BF01390325
  • [HPV15] Manfred Hartl, Teimuraz Pirashvili and Christine Vespa “Polynomial functors from algebras over a set-operad and nonlinear Mackey functors” In International Mathematics Research Notices 2015.6 Oxford University Press, 2015, pp. 1461–1554 DOI: 10.1093/imrn/rnt242
  • [Hô20] Quoc P Hô “Higher representation stability for ordered configuration spaces and twisted commutative factorization algebras”, 2020 arXiv:2004.00252v4
  • [Hô17] Quoc P. Hô “Free factorization algebras and homology of configuration spaces in algebraic geometry” In Selecta Mathematica 23.4 Springer, 2017, pp. 2437–2489 DOI: 10.1007/s00029-017-0339-1
  • [Hyd20] Trevor Hyde “Polynomial Factorization Statistics and Point Configurations in 3\mathbb{R}^{3} In International Mathematics Research Notices 2020.24 Oxford University Press, 2020, pp. 10154–10179 DOI: 10.1093/imrn/rny271
  • [Jim11] Rita Jiménez Rolland “Representation stability for the cohomology of the moduli space gn\mathcal{M}_{g}^{n} In Algebraic Geometry & Topology 11.5 Mathematical Sciences Publishers, 2011, pp. 3011–3041 DOI: 10.2140/agt.2011.11.3011
  • [Joy86] André Joyal “Foncteurs analytiques et espèces de structures” In Combinatoire énumérative Berlin, Heidelberg: Springer Berlin Heidelberg, 1986, pp. 126–159 DOI: 10.1007/BFb0072514
  • [Kri94] Igor Kriz “On the Rational Homotopy Type of Configuration Spaces” In Annals of Mathematics 139.2 Annals of Mathematics, 1994, pp. 227–237 DOI: 10.2307/2946581
  • [LS05] Riccardo Longoni and Paolo Salvatore “Configuration spaces are not homotopy invariant” In Topology 44.2 Elsevier, 2005, pp. 375–380 DOI: 10.1016/j.top.2004.11.002
  • [Mac95] Ian G. Macdonald “Symmetric functions and Hall polynomials” Oxford University Press, 1995
  • [Nie17] J. Nielsen “Die Isomorphismen der allgemeinen, unendlichen Gruppe mit zwei Erzeugenden” In Mathematische Annalen 78.1-4 Springer ScienceBusiness Media LLC, 1917, pp. 385–397 DOI: 10.1007/bf01457113
  • [Pag22] Roberto Pagaria “The Frobenius character of the Orlik-Terao algebra of type A”, 2022 arXiv:2203.08265v2
  • [Pet17] Dan Petersen “A spectral sequence for stratified spaces and configuration spaces of points” In Geometry & Topology 21.4 Mathematical Sciences Publishers, 2017, pp. 2527–2555 DOI: 10.2140/gt.2017.21.2527
  • [Pet20] Dan Petersen “Cohomology of generalized configuration spaces” In Compositio Mathematica 156.2 London Mathematical Society, 2020, pp. 251–298 DOI: 10.1112/S0010437X19007747
  • [PT] Dan Petersen and Orsola Tommasi “Multiplicative Chow–Künneth decomposition and homology splitting of configuration spaces” In preparation
  • [PT21] Dan Petersen and Phil Tosteson “Factorization statistics and bug-eyed configuration spaces” In Geometry & Topology 25.7 Mathematical Sciences Publishers, 2021, pp. 3691–3723 DOI: 10.2140/gt.2021.25.3691
  • [Pir00] Teimuraz Pirashvili “Hodge decomposition for higher order Hochschild homology” In Annales Scientifiques de l’École Normale Supérieure 33.2, 2000, pp. 151–179 Elsevier DOI: 10.1016/S0012-9593(00)00107-5
  • [Pow22] Geoffrey Powell “Baby bead representations”, 2022 arXiv:2209.08970v1
  • [Pow21] Geoffrey Powell “On analytic contravariant functors on free groups”, 2021 arXiv:2110.01934v1
  • [Pow22a] Geoffrey Powell “Outer functors and a general operadic framework”, 2022 arXiv:2201.13307v1
  • [PV18] Geoffrey Powell and Christine Vespa “Higher Hochschild homology and exponential functors”, 2018 arXiv:1802.07574v3
  • [Sta11] Richard P. Stanley “Enumerative Combinatorics” 1, Cambridge Studies in Advanced Mathematics Cambridge University Press, 2011 DOI: 10.1017/CBO9781139058520
  • [Sage9.3] The Sage Developers et al. “SageMath, version 9.3”, 2020 URL: http://www.sagemath.org
  • [Tot96] Burt Totaro “Configuration spaces of algebraic varieties” In Topology 35.4, 1996, pp. 1057–1067 DOI: https://doi.org/10.1016/0040-9383(95)00058-5
  • [TW17] Victor Turchin and Thomas Willwacher “Commutative hairy graphs and representations of Out(Fr)\operatorname{Out}(F_{r}) In Journal of Topology 10.2, 2017, pp. 386–411 DOI: 10.1112/topo.12009
  • [TW19] Victor Turchin and Thomas Willwacher “Hochschild-Pirashvili homology on suspensions and representations of Out(Fn)\operatorname{Out}({F}_{n}) In Annales Scientifiques de l’École Normale Supérieure 52.3, 2019, pp. 761–795 Société Mathématique de France (SMF) DOI: 10.24033/asens.2396
  • [Ves22] Christine Vespa “On the functors associated with beaded open Jacobi diagrams”, 2022 arXiv:2202.10907v1