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

Hopf algebras in the cohomology of 𝒜g\mathcal{A}_{g}, GLn()\mathrm{GL}_{n}(\mathbb{Z}), and SLn()\mathrm{SL}_{n}(\mathbb{Z})

Francis Brown All Souls College, Oxford, OX1 4AL, United Kingdom francis.brown@all-souls.ox.ac.uk Melody Chan Department of Mathematics, Brown University, Box 1917, Providence, RI 02912 melody_chan@brown.edu Søren Galatius Department of Mathematics, University of Copenhagen, Denmark galatius@math.ku.dk  and  Sam Payne Department of Mathematics, University of Texas at Austin, Austin, TX 78712 sampayne@utexas.edu
(Date: July 27, 2025)
Abstract.

We describe a bigraded cocommutative Hopf algebra structure on the weight zero compactly supported rational cohomology of the moduli space of principally polarized abelian varieties. By relating the primitives for the coproduct to graph cohomology, we deduce that dimHc2g+k(𝒜g)\dim H^{2g+k}_{c}(\mathcal{A}_{g}) grows at least exponentially with gg for k=0k=0 and for all but finitely many positive integers kk. Our proof relies on a new result of independent interest; we use a filtered variant of the Waldhausen construction to show that Quillen’s spectral sequence abutting to the cohomology of BK()BK(\mathbb{Z}) is a spectral sequence of Hopf algebras. From the same construction, we also deduce that dimH(n2)nk(SLn())\dim H^{\binom{n}{2}-n-k}(\mathrm{SL}_{n}(\mathbb{Z})) grows at least exponentially with nn, for k=1k=-1 and for all but finitely many non-negative integers kk.

1. Introduction

In this paper, we introduce and study new algebraic structures in the cohomology of the complex moduli space 𝒜=g0𝒜g\mathcal{A}=\bigsqcup_{g\geq 0}\mathcal{A}_{g} of principally polarized abelian varieties of all dimensions g0g\geq 0. Except where stated otherwise, all cohomology and compactly supported cohomology groups will be taken with coefficients in \mathbb{Q}.

The proofs of our results on the cohomology of 𝒜\mathcal{A} involve several technical constructions of independent interest. We produce a filtered coproduct on the Waldhausen construction of BK()BK(\mathbb{Z}), the de-looping of the KK-theory space of \mathbb{Z}. We also produce a filtered coproduct on a cubical space of graphs and a filtered map to BK()BK(\mathbb{Z}) that respects the relevant structures, inducing a morphism of spectral sequences of bialgebras. From the E1E^{1}-pages, we obtain new relations between the homology of the commutative graph complex 𝖦𝖢2\mathsf{GC}_{2} and K()K(\mathbb{Z}).

The terms on the E1E^{1}-page of Quillen’s spectral sequence, induced by the rank filtration on BK()BK(\mathbb{Z}), have natural interpretations in terms of the cohomology of GLn()\mathrm{GL}_{n}(\mathbb{Z}) and SLn()\mathrm{SL}_{n}(\mathbb{Z}). Thus, our constructions give new structures on the unstable cohomology of these groups.

1.1. A Hopf algebra structure on weight zero cohomology

The compactly supported cohomology Hc(𝒜)H^{*}_{c}(\mathcal{A}) is isomorphic to gHck(𝒜g)\bigoplus_{g}H^{k}_{c}(\mathcal{A}_{g}), which we consider as a bigraded \mathbb{Q}-vector space, in which Hck(𝒜g)H^{k}_{c}(\mathcal{A}_{g}) has bidegree (g,kg)(g,k\!-\!g). It inherits a natural mixed Hodge structure from those on Hc(𝒜g)H^{*}_{c}(\mathcal{A}_{g}), and the new structures we study are in its weight 0 subspace.

Theorem 1.1.

There is a bigraded Hopf algebra structure on the weight zero subspace W0Hc(𝒜)W_{0}H^{*}_{c}(\mathcal{A}).

The coproduct in this Hopf algebra structure is graded-cocommutative, and is induced by the proper maps 𝒜g×𝒜g𝒜g+g\mathcal{A}_{g}\times\mathcal{A}_{g^{\prime}}\to\mathcal{A}_{g+g^{\prime}} via pullback. The product we construct is more subtle; in particular, it is not graded-commutative.

We now identify a bigraded subspace of the primitives for the coproduct that is closely related to invariant differential forms on symmetric spaces. Let Ωc\Omega^{*}_{c} denote the vector space over \mathbb{Q} spanned by non-trivial exterior products of symbols ω4k+1\omega^{4k+1} for k1k\geq 1. It is bigraded by genus and degree minus genus, where ω4k1+1ω4kr+1\omega^{4k_{1}+1}\wedge\cdots\wedge\omega^{4k_{r}+1}, with k1<<krk_{1}<\cdots<k_{r}, has genus 2kr+12k_{r}+1 and degree (4k1+1)++(4kr+1)(4k_{1}+1)+\cdots+(4k_{r}+1). Let Ωc[1]\Omega^{*}_{c}[-1] denote the shift of Ωc\Omega^{*}_{c} in which ω4k1+1ω4kr+1\omega^{4k_{1}+1}\wedge\cdots\wedge\omega^{4k_{r}+1} has genus 2kr+12k_{r}+1 and degree (4k1+1)++(4kr+1)+1(4k_{1}+1)+\cdots+(4k_{r}+1)+1. Let Prim(H)\operatorname{Prim}(H) denote the subspace of primitives in a Hopf algebra HH.

Theorem 1.2.

There is an injection of bigraded vector spaces

Ωc[1]Prim(W0Hc(𝒜;)).\Omega^{*}_{c}[-1]\otimes\mathbb{R}\to\operatorname{Prim}(W_{0}H^{*}_{c}(\mathcal{A};\mathbb{R})).

This injection uses the identification W0Hc(𝒜g)Hc(Agtrop)W_{0}H^{*}_{c}(\mathcal{A}_{g})\cong H^{*}_{c}(A_{g}^{\mathrm{trop}}), where AgtropA_{g}^{\mathrm{trop}} is the moduli space of principally polarized tropical abelian varieties [BBC+24, BMV11], the stratification by locally symmetric spaces Agtrop=ggPg/GLg()A_{g}^{\mathrm{trop}}=\bigsqcup_{g^{\prime}\leq g}P_{g^{\prime}}/\mathrm{GL}_{g^{\prime}}(\mathbb{Z}), and the inclusion of the genus gg subspace of Ωc[1]\Omega^{*}_{c}[-1] in the compactly supported cohomology of Pg/GLg()P_{g}/\mathrm{GL}_{g}(\mathbb{Z}), as in [Bro23]. Here PgP_{g} denotes the cone of positive definite symmetric bilinear forms on g\mathbb{R}^{g}.

The subspace of primitives in a graded Hopf algebra is a Lie algebra with respect to the bracket [x,y]=xy(1)deg(x)deg(y)yx.[x,y]=xy-(-1)^{\deg(x)\deg(y)}yx. By relating this Lie algebra to the cohomology of the commutative graph complex 𝖦𝖢2\mathsf{GC}_{2} and the Grothendieck–Teichmüller Lie algebra, we find a subspace of Ωc[1]\Omega^{*}_{c}[-1], in diagonal bidegree (d,d)(d,d), that generates a free Lie subalgebra.

Theorem 1.3.

For N=11N=11, the image of {ω5,,ω4N+1}\{\omega^{5},\ldots,\omega^{4N+1}\} in Prim(W0Hc(𝒜;))\operatorname{Prim}(W_{0}H^{*}_{c}(\mathcal{A};\mathbb{R})) generates a free Lie subalgebra.

The statement for N=11N=11 reflects what we can prove with the current knowledge of the Grothendieck–Teichmüller Lie algebra; it is expected to hold for all NN.

By the Milnor–Moore theorem ([MM65, Theorem 5.18]), the cocommutative Hopf algebra W0Hc(𝒜)W_{0}H^{*}_{c}(\mathcal{A}) is isomorphic to 𝒰(Prim(W0Hc(𝒜)))\mathcal{U}(\operatorname{Prim}(W_{0}H^{*}_{c}(\mathcal{A}))), the universal enveloping algebra of its Lie algebra of primitives. The universal enveloping algebra on the free Lie algebra generated by the images of ω5,,ω4N+1\omega^{5},\ldots,\omega^{4N+1} is isomorphic to a tensor algebra generated by those elements. Therefore Theorems 1.2 and 1.3 have the following implications for dimensions. Let 𝒯\mathcal{T}_{*} denote a free genus-graded tensor algebra with one generator in each genus 3,5,,233,5,\ldots,23.

Corollary 1.4.

The dimension of W0Hc2g(𝒜g)W_{0}H^{2g}_{c}(\mathcal{A}_{g}) is greater than or equal to the dimension of 𝒯g\mathcal{T}_{g}. More generally, whenever Ωc[1]\Omega^{*}_{c}[-1] is nonzero in bidegree (g0,g0+k)(g_{0},g_{0}+k) for some nonnegative integer kk, the dimension of Hc2g+k(𝒜g)H^{2g+k}_{c}(\mathcal{A}_{g}) is greater than or equal to the dimension of 𝒯gg0\mathcal{T}_{g-g_{0}}.

Corollary 1.5.

The dimension of W0Hc2g+k(𝒜g)W_{0}H^{2g+k}_{c}(\mathcal{A}_{g}) grows at least exponentially with gg for k=0k=0 and all but finitely many positive integers kk.

See Section 7.3 for proofs and refinements of Corollaries 1.4 and 1.5. In particular, there are at most 20 exceptional values of kk for which dimHc2g+k(𝒜g)\dim H^{2g+k}_{c}(\mathcal{A}_{g}) does not grow at least exponentially. The precise dimension bounds given by Corollary 1.4 depend on knowing that the image of {ω5,,ω4N+1}\{\omega^{5},\ldots,\omega^{4N+1}\} generates a free Lie subalgebra for N=11N=11. However, the exponential growth stated in Corollary 1.5 already follows from the statement for N=2N=2. Extending Theorem 1.3 to larger values of NN improves the base of the exponential lower bound only slightly; see Section 7.2.


The complex moduli space 𝒜g\mathcal{A}_{g} is a smooth Deligne–Mumford stack and has the homotopy type of a classifying space for Sp2g()\mathrm{Sp}_{2g}(\mathbb{Z}). Thus, each of the results stated above has equivalent reformulations in terms of the group cohomology of Sp2g()\mathrm{Sp}_{2g}(\mathbb{Z}). Indeed, applying Poincaré duality and identifying the singular cohomology of 𝒜g\mathcal{A}_{g} with the group cohomology of Sp2g()\mathrm{Sp}_{2g}(\mathbb{Z}) gives

Hc(𝒜g)Hg2+g(Sp2g();).H^{*}_{c}(\mathcal{A}_{g})\cong H^{g^{2}+g-*}(\mathrm{Sp}_{2g}(\mathbb{Z});\mathbb{Q})^{\vee}.

In particular, we have the following immediate consequence of Corollary 1.5.

Corollary 1.6.

For k=0k=0 and all but finitely many positive integers kk, the dimension of Hg2gk(Sp2g();)H^{g^{2}-g-k}(\mathrm{Sp}_{2g}(\mathbb{Z});\mathbb{Q}) grows at least exponentially with gg.

It was previously known that Hg2g(𝒜g)H^{g^{2}-g}(\mathcal{A}_{g}) is nonvanishing for all gg, because the tautological subring of H(𝒜g)H^{*}(\mathcal{A}_{g}), i.e., the subring generated by the λ\lambda-classes, is Gorenstein with socle in this degree [vdG99]. It is expected that Hk(𝒜g)H^{k}(\mathcal{A}_{g}) may vanish for k>g2gk>g^{2}-g [BPS23, Question 1.1]. We note that all classes in the tautological subring are of pure weight, i.e., weight equal to cohomological degree, and the tautological subspace of Hg2g(𝒜g)H^{g^{2}-g}(\mathcal{A}_{g}) always has rank 1. The growth we find is in the top weight cohomology, i.e., the graded piece of the weight filtration Poincaré-dual to the weight zero compactly supported cohomology.

1.2. Filtered coproduct on the Waldhausen construction

The heart of our proof of Theorems 1.1 and 1.2 is the construction of a coassociative but not cocommutative coproduct on the Waldhausen construction of BK()BK(\mathbb{Z}) [Wal85] that is compatible with the rank filtration. Here, BK()BK(\mathbb{Z}) is the 1-fold delooping of the KK-theory space for \mathbb{Z}, so πi+1(BK())=Ki()\pi_{i+1}(BK(\mathbb{Z}))=K_{i}(\mathbb{Z}). The homological spectral sequence associated to the rank filtration on BK()BK(\mathbb{Z}) is called the Quillen spectral sequence, and we denote it by EQ{}^{Q}\!E^{*}. The terms on the E1E^{1}-page satisfy

(1) En,k1QHk(GLn();Stn),{}^{Q}\!E^{1}_{n,k}\cong H_{k}(\mathrm{GL}_{n}(\mathbb{Z});\operatorname{St}_{n}\otimes\mathbb{Q}),

where Stn\operatorname{St}_{n} denotes the Steinberg module. Our filtered coproduct gives rise to a bigraded Hopf algebra structure on each page of EQ{}^{Q}\!E^{*}, as in Theorem 3.18. Let Indec(H)\mathrm{Indec}(H) denote the indecomposables in a Hopf algebra HH. For E1Q{}^{Q}\!E^{1} we have the following.

Theorem 1.7.

There is a bigraded commutative Hopf algebra structure on the E1E^{1}-page of the Quillen spectral sequence

En,k1Q=Hk(GLn();Stn),{}^{Q}\!E^{1}_{n,k}=H_{k}(\mathrm{GL}_{n}(\mathbb{Z});\operatorname{St}_{n}\otimes\mathbb{Q}),

and a surjection of bigraded vector spaces

(2) Indec(E1Q)𝜋Ωc[1]e.\mathrm{Indec}({}^{Q}\!E^{1}\otimes\mathbb{R})\xrightarrow{\pi}\Omega^{*}_{c}[-1]^{\vee}\oplus\mathbb{R}\cdot e.

Here, ee is in bidegree (1,0)(1,0).

These results follow from Theorem 3.18 and the proof of Theorem 1.2 in Section 4 (establishing an injection dual to (2)).

1.3. Relation to the graph spectral sequence

Our proof of Theorem 1.3 relies on relations between the filtered coproduct on the Waldhausen construction of BK()BK(\mathbb{Z}) and an analogous filtered coproduct on a space of graphs BKGrBK_{\mathrm{Gr}}. This filtered space of graphs, which will be defined in Section 5, has the rational homology of the disjoint union of a circle and a point, but the graded dual of the E1E_{1}-page of the associated cohomological spectral sequence is a bialgebra whose primitive elements contain a copy of the Grothendieck–Teichmüller Lie algebra.

Moreover, the space BKGrBK_{\mathrm{Gr}} admits a filtered map to the Waldhausen construction of BK()BK(\mathbb{Z}) that respects all of the relevant structures. Let Er,Q{}^{Q}\!E_{r}^{*,*} denote the pages of the cohomological spectral sequence dual to E,rQ{}^{Q}\!E^{r}_{*,*}. The induced map on the diagonal subspace E1g,gE_{1}^{g,g} will be most important.

Proposition 1.8.

There is a morphism of graded Lie algebras

(3) Prim(gE1g,gQ)H0(𝖦𝖢2)\operatorname{Prim}\Big{(}\bigoplus_{g}{}^{Q}\!E_{1}^{g,g}\Big{)}\to H^{0}(\mathsf{GC}_{2})

which, after tensoring with \mathbb{R}, sends ω4k+1\omega^{4k+1} for all k1k\geq 1 to an element that is non-trivial in the abelianization of H0(𝖦𝖢2)H^{0}(\mathsf{GC}_{2})\otimes\mathbb{R} with respect to its graded Lie algebra structure.

Moreover, for N=11N=11, there is an injection

T(ω5,ω9,,ω4N+1)g0E1g,gQ.T(\langle\omega^{5},\omega^{9},\ldots,\omega^{4N+1}\rangle)\hookrightarrow\bigoplus_{g\geq 0}{}^{Q}\!E_{1}^{g,g}\otimes\mathbb{R}.

This result is established in (82), Proposition 6.2, and Corollary 6.3. The morphism (3) is induced by the tropical Torelli map, up to a zig-zag of rationally equivalent spaces that is explained in Section 2.4 and further in Section 6. The class ω4k+1\omega^{4k+1} pairs non-trivially with the homology class of the wheel graph [W2k+1]H0(𝖦𝖢2)[W_{2k+1}]\in H_{0}(\mathsf{GC}_{2}^{\vee}). This pairing is given by an integral which is a non-zero rational multiple of ζ(2k+1)\zeta(2k+1) [BS24]. To prove Theorem 1.3, we use known results on the Grothendieck–Teichmüller Lie algebra, which is isomorphic to H0(𝖦𝖢2)H^{0}(\mathsf{GC}_{2}) by [Wil15], the fact that it contains a free Lie algebra with one generator in each odd degree greater than one [Bro12], and best-known bounds on agreement between the two with respect to their weight filtrations. See Section 6.1.

We may pass from such statements about EQ{}^{Q}\!E_{*} to statements on W0Hc(𝒜g)W_{0}H^{*}_{c}(\mathcal{A}_{g}). Using one of the main results of [BBC+24], we show that there is a canonical choice of an element ee in E1,01Q{}^{Q}\!E^{1}_{1,0} and an isomorphism of bigraded algebras

(4) (W0Hc(𝒜))[x]/(x2)E,1Q(W_{0}H^{*}_{c}(\mathcal{A}))^{\vee}\otimes_{\mathbb{Q}}\mathbb{Q}[x]/(x^{2})\overset{\sim}{\to}{}^{Q}\!E^{1}_{*,*}

taking xx to ee. The element ee is a primitive in the Hopf algebra structure on E,1Q{}^{Q}\!E^{1}_{*,*}, and Theorem 1.1 is proved by taking the quotient by the Hopf ideal generated by this primitive, and passing to the graded dual. See Section 4.

1.4. Applications to the unstable cohomology of GLn()\mathrm{GL}_{n}(\mathbb{Z}) and SLn()\mathrm{SL}_{n}(\mathbb{Z})

Our construction of a Hopf algebra structure on Quillen’s spectral sequence, together with the injection in Theorem 1.2, gives the following result and produces a bigraded commutative Hopf algebra structure on a large subspace of nH(SLn();)\bigoplus_{n}H^{*}(\mathrm{SL}_{n}(\mathbb{Z});\mathbb{Q}).

Corollary 1.9.

The dimensions of H(n2)nk(GLn();or)H^{\binom{n}{2}-n-k}(\mathrm{GL}_{n}(\mathbb{Z});\mathbb{Q}_{\mathrm{or}}) and of H(n2)nk(SLn();)H^{\binom{n}{2}-n-k}(\mathrm{SL}_{n}(\mathbb{Z});\mathbb{Q}) grow at least exponentially with nn for k=1k=-1 and for all but finitely many nonnegative integers kk.

See Section 7.3.2, in particular for the refined statement that there are at most 10 exceptional values of kk for which these groups do not grow at least exponentially.

The cohomology groups Hi(SLn();)H^{i}(\mathrm{SL}_{n}(\mathbb{Z});\mathbb{Q}) are stable in low degrees, for in2i\leq n-2 [LS19]; the growth in Corollary 1.9 takes place entirely in the unstable range. In the highest degrees, H(n2)nk(SLn();)H^{\binom{n}{2}-n-k}(\mathrm{SL}_{n}(\mathbb{Z});\mathbb{Q}) is conjectured to vanish for k<1k<-1 [CFP14, Conjecture 2]. Hence, Corollary 1.9 says that the cohomology groups grow exponentially in the degree immediately below the threshold at and above which they are expected to vanish, and at almost every fixed distance below the threshold.

The following corollary can be deduced from dualizing (2) in Theorem 1.7, together with the Poincaré–Birkhoff–Witt theorem and Shapiro’s Lemma.

Corollary 1.10.

There is an injection of bigraded vector spaces

Sym(Ωc[1]ϵ)nH(n+12)(SLn();)\operatorname{Sym}\left(\Omega^{*}_{c}[-1]\oplus\mathbb{Q}\cdot\epsilon\right)\otimes\mathbb{R}\hookrightarrow\bigoplus_{n}H_{\binom{n+1}{2}-*}(\mathrm{SL}_{n}(\mathbb{Z});\mathbb{R})

Here, ϵ\epsilon is in bidegree (1,0)(1,0), and H(n+12)k(SLn())H_{\binom{n+1}{2}-k}(\mathrm{SL}_{n}(\mathbb{Z})) is in bidegree (n,kn)(n,k-n).

Here, and elsewhere, Sym\operatorname{Sym} denotes the free graded-commutative algebra on a graded vector space. It is the tensor product of an exterior algebra on elements of Ωc[1]ϵ\Omega^{*}_{c}[-1]\oplus\mathbb{R}\epsilon of odd degree, and a commutative polynomial algebra on the elements of even degree. A version of this statement (without the class ϵ\epsilon) was announced by Ronnie Lee [Lee78], but no proof has appeared in the literature. Our results imply that the symmetric algebra on a much larger set of primitive elements embeds into the cohomology of the special linear group: in addition to the classes in Ωc[1]ϵ\Omega^{*}_{c}[-1]\oplus\mathbb{Q}\cdot\epsilon, we may also take infinitely many independent commutators in the free Lie algebra on {ω5,,ω45}\{\omega^{5},\ldots,\omega^{45}\}; see Remark 7.3.

The injection in Corollary 1.10 is defined with real coefficients. We also give related constructions with rational coefficients. The wheel graphs W2k+1W_{2k+1}, for positive integers kk, give non-trivial homology classes in the graph homology H0(𝖦𝖢2)H_{0}(\mathsf{GC}_{2}^{\vee}), i.e., the graded dual of H0(𝖦𝖢2)H^{0}(\mathsf{GC}_{2}). Moreover, [W2k+1][W_{2k+1}] pairs nontrivially with ω4k+1,\omega^{4k+1}, by pushing forward from the filtered space of graphs BKGrBK_{\mathrm{Gr}} discussed in Section 1.3 and identifying [ω4k+1][\omega^{4k+1}] with a compactly supported cohomology class on the rank-(2k+1)(2k+1) stratum of BK()BK(\mathbb{Z}). By showing that these odd wheel classes are primitive with respect to the coproduct, we deduce the following statement in Section 7.4.

Corollary 1.11.

There is an injective map of commutative bigraded algebras

[W3,W5,,W2k+1,][x]/(x2)E,1Q\mathbb{Q}[W_{3},W_{5},\ldots,W_{2k+1},\ldots]\otimes\mathbb{Q}[x]/(x^{2})\longrightarrow{}^{Q}\!E^{1}_{*,*}

where the element xx is in bidegree (1,0)(1,0).

Remark 1.12.

Note that the last two corollaries only give polynomial growth in the cohomology of 𝒜\mathcal{A} and SLn()\mathrm{SL}_{n}(\mathbb{Z}). Our results on exponential growth are stronger and fundamentally use the fact that our co-commutative Hopf algebras are far from commutative.

Remark 1.13.

When this paper was in the final stages of preparation, we learned that Ash, Miller, and Patzt independently discovered a bigraded commutative Hopf algebra structure on k,nHk(GLn();Stn)\bigoplus_{k,n}H_{k}(\mathrm{GL}_{n}(\mathbb{Z});\operatorname{St}_{n}\otimes\mathbb{Q}) and described some indecomposable elements. It is not immediately clear whether the two Hopf algebra structures agree. They also found a related Hopf algebra structure on k,nHk(GLn();Stn~)\bigoplus_{k,n}H_{k}(\mathrm{GL}_{n}(\mathbb{Z});\operatorname{St}_{n}\otimes\widetilde{\mathbb{Q}}) [AMP24].

Remark 1.14.

Recent work of Yuan and Jansen introduces an EE_{\infty}-monoid denoted K()K^{\partial}(\mathbb{Z}) in [Yua23] and |RBS()||\mathcal{M}_{\mathrm{RBS}}(\mathbb{Z})| in [Jan24] which seems related to the Hopf algebra structure on the Quillen spectral sequence, presumably by Koszul duality. In particular, [Jan24, Theorem 10.11] implies that the E1E^{1}-page of Quillen’s spectral sequence arises as homology of the bar construction of an EE_{\infty}-algebra, which can perhaps be viewed as a conceptual reason for the Hopf algebra structure.

1.5. Relations to moduli spaces of curves

Theorems 1.1 and 1.3, and the resulting exponential growth of dimW0Hc2g(𝒜g)\dim W_{0}H^{2g}_{c}(\mathcal{A}_{g}) (Corollary 1.5 for k=0k=0) are analogs of the main results of [CGP21] for the moduli space of curves =g2g\mathcal{M}=\bigsqcup_{g\geq 2}\mathcal{M}_{g}. Neither of these collections of results appears to be directly deducible from the other. However, W0Hc(𝒜)W_{0}H^{*}_{c}(\mathcal{A}) and W0Hc()W_{0}H^{*}_{c}(\mathcal{M}) can be related to each other through a zig-zag, as follows. Recall the Torelli map 𝒜\mathcal{M}\to\mathcal{A}, taking a curve to its Jacobian. The Torelli map is not proper, and hence does not give rise to a natural morphism of mixed Hodge structures between Hc()H^{*}_{c}(\mathcal{M}) and Hc(𝒜)H^{*}_{c}(\mathcal{A}), but it factors as ct𝒜\mathcal{M}\to\mathcal{M}^{\mathrm{ct}}\to\mathcal{A}, through the open inclusion of \mathcal{M} into the moduli space ct\mathcal{M}^{\mathrm{ct}} of curves of compact type (or the image of ct\mathcal{M}^{\mathrm{ct}} in 𝒜\mathcal{A}). The Torelli map extends to a proper morphism on ct\mathcal{M}^{\mathrm{ct}}, giving rise to a zig-zag of mixed Hodge structures

Hc()Hc(ct)Hc(𝒜).H^{*}_{c}(\mathcal{M})\to H^{*}_{c}(\mathcal{M}^{\mathrm{ct}})\leftarrow H^{*}_{c}(\mathcal{A}).

However, we do not yet have sufficient understanding of the compactly supported cohomology of ct\mathcal{M}^{\mathrm{ct}} (or its image in 𝒜\mathcal{A}) to use such a zig-zag effectively.

Nevertheless, the Grothendieck–Teichmüller Lie algebra, via its interpretation as the zeroth degree cohomology of the commutative graph complex 𝖦𝖢2\mathsf{GC}_{2}, plays a common role in the top weight compactly supported cohomology of both 𝒜\mathcal{A} and \mathcal{M}. Indeed, the main results of [CGP21] identify W0Hc()W_{0}H^{*}_{c}(\mathcal{M}) with H(𝖦𝖢2)H^{*}(\mathsf{GC}_{2}) and thereby endow it with the structure of a bigraded Lie algebra. Combining this identification with results from Grothendieck–Teichmüller theory [Bro12, Wil15] shows that gW0Hc2g(g)\bigoplus_{g}W_{0}H^{2g}_{c}(\mathcal{M}_{g}) contains a free Lie subalgebra with one generator σg\sigma_{g} in each odd genus g3g\geq 3 and hence dimW0Hc2g(g)\dim W_{0}H^{2g}_{c}(\mathcal{M}_{g}) grows at least exponentially with gg. In particular the graph cohomology H(𝖦𝖢2)H^{*}(\mathsf{GC}_{2}) plays the role of an intermediary between W0Hc()W_{0}H^{*}_{c}(\mathcal{M}) and W0Hc(𝒜)W_{0}H^{*}_{c}(\mathcal{A}). Indeed, the main construction of [CGP21] is a map from 𝖦𝖢2\mathsf{GC}_{2} to a cellular chain complex that computes W0Hc()W_{0}H^{*}_{c}(\mathcal{M}), which is in fact a quasi-isomorphism. Here, we construct a map from Prim(E1Q)\operatorname{Prim}\big{(}{}^{Q}\!E_{1}\big{)} to H(𝖦𝖢2)H^{*}(\mathsf{GC}_{2}) that vanishes on E11,0Q{}^{Q}\!E_{1}^{1,0} and hence factors as

Prim(E1Q)Prim(W0Hc(𝒜))H(𝖦𝖢2).\operatorname{Prim}\big{(}{}^{Q}\!E_{1}\big{)}\to\operatorname{Prim}\big{(}W_{0}H^{*}_{c}(\mathcal{A})\big{)}\to H^{*}(\mathsf{GC}_{2}).

This map Prim(W0Hc(𝒜))H(𝖦𝖢2)\operatorname{Prim}\big{(}W_{0}H^{*}_{c}(\mathcal{A})\big{)}\to H^{*}(\mathsf{GC}_{2}) is not an isomorphism, but is nevertheless essential to our proof of Theorem 1.3 and its corollaries. It remains possible that this map may restrict to an isomorphism Prim(gHc2g(𝒜))H0(𝖦𝖢2)\operatorname{Prim}\big{(}\bigoplus_{g}H^{2g}_{c}(\mathcal{A})\big{)}\to H^{0}(\mathsf{GC}_{2}); see Question 1.17.

Remark 1.15.

The analog of Corollary 1.5 for Hc()H^{*}_{c}(\mathcal{M}) is an open problem; it is conjectured but not known that dimHc2g+k(g)\dim H^{2g+k}_{c}(\mathcal{M}_{g}) grows at least exponentially with gg for all but finitely many k0k\geq 0 [PW24, Conjecture 1.3]. This is proved for a few dozen values of kk. In each known case, the exponential growth is established in one fixed graded piece of the weight filtration; see [PW21, Corollary 1.3] and [PW24, Corollary 1.2].

1.6. Further results and conjectures

We now state questions, conjectures, and further results related to injectivity, vanishing, and extensions of Hodge structures involving W0Hc(𝒜)W_{0}H^{*}_{c}(\mathcal{A}). We also state a generalization of our results related to the Hopf algebra on the Quillen spectral sequence (Theorem 1.7 and Proposition 1.8) for rings of integers in number fields.

1.6.1. Injectivity

Theorems 1.2 and 1.3 suggest the following conjecture.

Conjecture 1.16.

The inclusion of Ωc[1]\Omega^{*}_{c}[-1]\otimes\mathbb{R} into the primitives for the coproduct on W0Hc(𝒜;)W_{0}H^{*}_{c}(\mathcal{A};\mathbb{R}) induces an injection T(Ωc[1])W0Hc(𝒜;).T(\Omega^{*}_{c}[-1])\otimes\mathbb{R}\to W_{0}H^{*}_{c}(\mathcal{A};\mathbb{R}).

To support this conjecture, we construct a spectral sequence of bigraded Hopf algebras whose E1E_{1}-page is T(Ωc[1])T(\Omega^{*}_{c}[-1]) and show that the abutment of this spectral sequence is isomorphic, as a bigraded vector space, to the abutment of the Quillen spectral sequence. Note that Conjecture 1.16 implies in particular that the tensor algebra on {ω4k+1}\{\omega^{4k+1}\} injects into W0Hc(𝒜;)W_{0}H^{*}_{c}(\mathcal{A};\mathbb{R}), or equivalently, that the classes ω4k+1\omega^{4k+1} generate a free Lie subalgebra of Prim(W0Hc(𝒜;))\operatorname{Prim}(W_{0}H^{*}_{c}(\mathcal{A};\mathbb{R})). To prove this weaker statement, it would suffice to show that each class ω4k+1\omega^{4k+1} maps into the motivic Lie subalgebra of the Grothendieck–Teichmüller Lie algebra [Bro12].

The map in Conjecture 1.16 is injective for g9g\leq 9 and an isomorphism for g7g\leq 7; the former follows from the methods developed in this paper, the latter from the isomorphism (4) and explicit calculations in [EVGS13]. It is not expected to be an isomorphism in general. Indeed, the dual vector space of Ωc[1]\Omega^{*}_{c}[-1] in genus gg is expected to be spanned by cohomology classes for GLg()\mathrm{GL}_{g}(\mathbb{Z}) of non-cuspidal (Eisenstein) type [Gro13]. Using automorphic methods, cuspidal cohomology classes for GLg()\mathrm{GL}_{g}(\mathbb{Z}) for g=79g=79 and 105105 were recently constructed in [BCG23]. Also, starting from g=9g=9, the Euler characteristic of T(Ωc[1])T(\Omega^{*}_{c}[-1]) and the isomorphism (4) are not compatible with homological Euler characteristic computations for GLg()\mathrm{GL}_{g}(\mathbb{Z}) [Hor05, Theorem 3.3].

The situation potentially changes if we restrict to the diagonal gW0Hc2g(𝒜g;)\bigoplus_{g}W_{0}H^{2g}_{c}(\mathcal{A}_{g};\mathbb{R}), which, by our earlier results, is a graded cocommutative Hopf algebra of particular interest.

Question 1.17.

Is the induced map

T(k1ω4k+1)gW0Hc2g(𝒜g;)T\Big{(}\bigoplus_{k\geq 1}\omega^{4k+1}\mathbb{R}\Big{)}\to\bigoplus_{g}W_{0}H^{2g}_{c}(\mathcal{A}_{g};\mathbb{R})

an isomorphism?

As noted above, the injectivity of this map would follow from Drinfeld’s conjecture that the motivic Lie algebra surjects onto the Grothendieck–Teichmüller Lie algebra. Assuming the same conjecture and restricting to the Lie algebra of primitives, the analog of Question 1.17 for the cohomology of the moduli stack of curves has an affirmative answer, since gW0Hc2g(g)\bigoplus_{g}W_{0}H^{2g}_{c}(\mathcal{M}_{g}) is isomorphic to H0(𝖦𝖢2)H^{0}(\mathsf{GC}_{2}).

1.6.2. Vanishing

Note that Hci(g)H^{i}_{c}(\mathcal{M}_{g}) vanishes for i<2gi<2g [Har86, CFP12, MSS13]. Passing to weight zero, this implies the vanishing of H(𝖦𝖢2)H^{*}(\mathsf{GC}_{2}) in negative degrees; see [Wil15, Theorem 1.1] and [CGP21, Theorem 1.4]. It is expected that Hci(𝒜g)H^{i}_{c}(\mathcal{A}_{g}) may also vanish for i<2gi<2g [BPS23, Question 1.1]; this is known for i<g+max{2,g}i<g+\max\{2,g\} [BS73, Gun00, BPS23].

The vanishing of Hc2g+1(g)H^{2g+1}_{c}(\mathcal{M}_{g}) for all gg is a compelling open question, closely related to results in deformation theory, such as formality of deformation quantization [PW24, Question 1.4]. The analogous statement for Hc(𝒜)H^{*}_{c}(\mathcal{A}) is also open.

Question 1.18.

Does Hc2g+1(𝒜g)H^{2g+1}_{c}(\mathcal{A}_{g}) vanish for all gg?

The answer is yes for g4g\leq 4. See [Hai02] and [HT12, Corollary 3].

1.6.3. Extensions of Tate Hodge structures

Our study of W0Hc(𝒜)W_{0}H^{*}_{c}(\mathcal{A}) also sheds new light on the full mixed Hodge structure Hc(𝒜)H^{*}_{c}(\mathcal{A}). In particular, we can now show that this mixed Hodge structure contains a nontrivial extension of Tate Hodge structures in genus 3, answering questions of Hain and Looijenga.

Following Namikawa [Nam80], we shall use the notation 𝒜g\mathcal{A}_{g}^{\textrm{\begin{CJK}{UTF8}{min}サ\end{CJK}}} for the Satake–Baily–Borel compactification of 𝒜g\mathcal{A}_{g}. Hain observed over twenty years ago that the mixed Hodge structure on H6(𝒜3)H^{6}(\mathcal{A}_{3}) is an extension of (6)\mathbb{Q}(-6) by (3)\mathbb{Q}(-3) and stated the expectation that it should be a multiple (possibly trivial) of the extension given by ζ(3)\zeta(3) [Hai02, pp. 473–474]. More recently, Looijenga showed that a Tate twist of a nontrivial multiple of this same extension appears in the stable cohomology of the Satake compactification H6(𝒜)H^{6}(\mathcal{A}_{\infty}^{\textrm{\begin{CJK}{UTF8}{min}サ\end{CJK}}}) and asked whether the pullback H6(𝒜)H6(𝒜3)H^{6}(\mathcal{A}_{\infty}^{\textrm{\begin{CJK}{UTF8}{min}サ\end{CJK}}})\to H^{6}(\mathcal{A}_{3}^{\textrm{\begin{CJK}{UTF8}{min}サ\end{CJK}}}) is injective [Loo17, p. 1370]. Here, we confirm Hain’s expectation and give an affirmative answer to Looijenga’s question. Moreover, we show that the extension in H6(𝒜3)H^{6}(\mathcal{A}_{3}) is nontrivial.

Theorem 1.19.

The restriction map H6(𝒜)H6(𝒜3)H^{6}(\mathcal{A}_{\infty}^{\textrm{\begin{CJK}{UTF8}{min}サ\end{CJK}}})\to H^{6}(\mathcal{A}_{3}^{\textrm{\begin{CJK}{UTF8}{min}サ\end{CJK}}}) is injective. Moreover, the image of this restriction map is equal to the image of Hc6(𝒜3)H^{6}_{c}(\mathcal{A}_{3}) under push forward for the open inclusion 𝒜3𝒜3\mathcal{A}_{3}\subset\mathcal{A}_{3}^{\textrm{\begin{CJK}{UTF8}{min}サ\end{CJK}}}. In particular, the mixed Hodge structure on Hc6(𝒜3)H^{6}_{c}(\mathcal{A}_{3}) is the nontrivial extension of (3)\mathbb{Q}(-3) by \mathbb{Q} given by a nonzero rational multiple of ζ(3)\zeta(3).

Theorem 1.19 is proved in Section 7.7. By Poincaré duality, we see that H6(𝒜3)H^{6}(\mathcal{A}_{3}) is the nontrivial extension of (6)\mathbb{Q}(-6) by (3)\mathbb{Q}(-3) given by a nonzero rational multiple of ζ(3)\zeta(3).

Remark 1.20.

The motivic structure of H6(𝒜3)H^{6}(\mathcal{A}_{3}) (i.e. its associated mixed Hodge structure and \ell-adic Galois representations) and the extensions of Tate structures in H2g(𝒜g)H^{2g}(\mathcal{A}_{g}^{\textrm{\begin{CJK}{UTF8}{min}サ\end{CJK}}}) have generated sustained interest. See, for instance, [CFvdG20, Section 13] for a discussion of the Siegel and Teichmüller modular forms that occur in H6(𝒜3)H^{6}(\mathcal{A}_{3}) and H6(3)H^{6}(\mathcal{M}_{3}) and [vdGL21] for a discussion of the differences that appear when working over fields of positive characteristic.

We predict that Theorem 1.19 generalizes to higher genera, as follows.

Conjecture 1.21.

For odd g3g\geq 3, the mixed Hodge structure on Hc2g(𝒜g)H^{2g}_{c}(\mathcal{A}_{g}) has a subquotient isomorphic to the extension of (g)\mathbb{Q}(-g) by \mathbb{Q} given by a nonzero rational multiple of ζ(g)\zeta(g).

As evidence for Conjecture 1.21, we note that GrW2gHc2g(𝒜g)\operatorname{Gr}_{W}^{2g}H^{2g}_{c}(\mathcal{A}_{g}) contains a copy of (g)\mathbb{Q}(-g), because the tautological subring of H(𝒜g)H^{*}(\mathcal{A}_{g}) has its socle in degree g2gg^{2}-g [vdG99].

1.7. Quillen spectral sequences for rings of integers in number fields

Our proof of the existence of a Hopf algebra structure on the Quillen spectral sequence in Theorem 1.7 holds for much more general rings RR. A particular case of interest is when R=𝒪KR=\mathcal{O}_{K} is the ring of integers in a number field KK. In this setting, its E1E^{1}-page was computed by Quillen and satisfies

En,k1Q(𝒪K)𝔞Hk(GL(P𝔞);St(P𝔞K)){}^{Q}\!E_{n,k}^{1}(\mathcal{O}_{K})\cong\bigoplus_{\mathfrak{a}}H_{k}(\mathrm{GL}(P_{\mathfrak{a}});\operatorname{St}(P_{\mathfrak{a}}\otimes K))

where the sum is over representatives P𝔞P_{\mathfrak{a}} of the isomorphism classes of projective 𝒪K\mathcal{O}_{K}-modules of rank nn (which for n>0n>0 are in bijection with the ideal class group of 𝒪K\mathcal{O}_{K}). Our results imply that E,rQ(𝒪K){}^{Q}\!E^{r}_{*,*}(\mathcal{O}_{K}), and in particular its E1E^{1}-page, has the structure of a commutative bigraded Hopf algebra. It gives a spectral sequence of Hopf algebras converging to the rational homology of BK(𝒪K)BK(\mathcal{O}_{K}), which was computed by Borel.


Acknowledgments. We are most grateful to Samuel Grushevsky for illuminating discussions about the geometry of 𝒜g\mathcal{A}_{g} and its compactifications. We also thank Avner Ash, Dick Hain, Ralph Kaufmann, Jeremy Miller, Natalia Pacheco-Tallaj, Peter Patzt, Dan Petersen, and Jan Steinebrunner for helpful conversations.

FB thanks Trinity College Dublin for a Simons Visiting Professorship, and the University of Geneva for hospitality during which part of this work was carried out. MC was supported by NSF CAREER DMS-1844768, NSF FRG DMS–2053221, and a grant from the Simons Foundation (1031656, Chan). SG thanks Alexander Kupers and Oscar Randal-Williams for previous collaborations on related topics [GKRW20], and was supported by the Danish National Research Foundation (DNRF151). Part of this work was carried out while SG held a one-year visiting position at Columbia University, and he thanks the department for its hospitality and support. SP was supported by NSF Grant DMS–2302475, NSF FRG Grant DMS–2053261, and a CRM–Simons Visiting Professorship in Montréal. Part of this collaboration occurred at Stanford University supported by the Poincaré Distinguished Visiting Professorship of SP.

2. Preliminaries

2.1. Weight zero cohomology and the tropical spectral sequence

Let 𝒜g\mathcal{A}_{g} denote the complex moduli stack of principally polarized abelian varieties of dimension gg. For background on geometry and topology of 𝒜g\mathcal{A}_{g} and its compactifications, we refer to the articles [Gru09] and [HT18]. For each k0k\geq 0, its compactly supported rational cohomology Hck(𝒜g;)H^{k}_{c}(\mathcal{A}_{g};\mathbb{Q}) admits a weight filtration

W0Hck(𝒜g;)W1Hck(𝒜g;)Hck(𝒜g;)W_{0}H^{k}_{c}(\mathcal{A}_{g};\mathbb{Q})\subset W_{1}H^{k}_{c}(\mathcal{A}_{g};\mathbb{Q})\subset\cdots\subset H^{k}_{c}(\mathcal{A}_{g};\mathbb{Q})

as part of the mixed Hodge structure on Hck(𝒜g;)H^{k}_{c}(\mathcal{A}_{g};\mathbb{Q}). Let 𝒜=g0𝒜g\mathcal{A}=\coprod_{g\geq 0}\mathcal{A}_{g} be the moduli space of principally polarized abelian varieties of any dimension. Then Hc(𝒜)=g0Hc(𝒜g)H^{*}_{c}(\mathcal{A})=\bigoplus_{g\geq 0}H^{*}_{c}(\mathcal{A}_{g}) and we take WkHc(𝒜)=gWkHc(𝒜g)W_{k}H^{*}_{c}(\mathcal{A})=\bigoplus_{g}W_{k}H^{*}_{c}(\mathcal{A}_{g}).

There is no single canonical choice of normal crossings compactification for 𝒜g\mathcal{A}_{g}. Rather, by [AMRT75], there is a toroidal compactification 𝒜g¯Σ\overline{\mathcal{A}_{g}}^{\Sigma} for every choice of certain polyhedral data Σ\Sigma, as we now recall. Let PgSym2((g))P_{g}\subset\mathrm{Sym}^{2}((\mathbb{R}^{g})^{\vee}) denote the set of symmetric bilinear forms on g\mathbb{R}^{g} that are positive definite. It is an open convex cone of full dimension inside the (g+12)\binom{g+1}{2}-dimensional Euclidean vector space Sym2((g))\mathrm{Sym}^{2}((\mathbb{R}^{g})^{\vee}). Let PgrtP_{g}^{\mathrm{rt}} denote the set of positive semidefinite forms on g\mathbb{R}^{g} whose kernel is rational, i.e., the kernel is of the form WW\otimes\mathbb{R} for a vector subspace WW of g\mathbb{Q}^{g}. Thus PgPgrtP_{g}\subset P_{g}^{\mathrm{rt}}. Let Σ\Sigma denote any admissible decomposition of PgrtP_{g}^{\mathrm{rt}}. That is, Σ\Sigma is an infinite rational polyhedral cone decomposition supported on PgrtP_{g}^{\mathrm{rt}}, whose cones are permuted under the action given by XA:=ATXAX\cdot A:=A^{T}XA for AGLg()A\in\mathrm{GL}_{g}(\mathbb{Z}), and there are only finitely many orbits of cones of Σ\Sigma under this action. It is a nontrivial, but classical, fact that admissible decompositions exist. Famous examples include the decomposition into perfect cones; see [AMRT75] and references therein.

It will be convenient to topologize PgrtP_{g}^{\mathrm{rt}} not with its subspace topology induced from the ambient vector space Sym2((g))\mathrm{Sym}^{2}((\mathbb{R}^{g})^{\vee}), but rather with its Satake topology, which we now define. Let Σ\Sigma denote any admissible decomposition of PgrtP_{g}^{\mathrm{rt}}. The Satake topology on the set PgrtP_{g}^{\mathrm{rt}} is the finest topology such that for every cone σ\sigma in Σ\Sigma, the map σPgrt\sigma\to P_{g}^{\mathrm{rt}} is continuous. The Satake topology agrees with the Euclidean topology on PgrtP_{g}^{\mathrm{rt}} for g=0,1g=0,1, but is strictly finer for g2g\geq 2. On the other hand, the Satake topology restricts to the Euclidean topology on the open subset PgP_{g}, since every point of PgP_{g} is contained in only finitely many cones of Σ\Sigma. Moreover, it is independent of the choice of Σ\Sigma, since any two choices of admissible decompositions admit a common refinement [FC90, IV.2, p. 97]. Throughout, we consider PgrtP_{g}^{\mathrm{rt}} as a topological space with the Satake topology.

Now define

Agtrop:=Pgrt/GLg().A_{g}^{\mathrm{trop}}:=P_{g}^{\mathrm{rt}}/\mathrm{GL}_{g}(\mathbb{Z}).

For an interpretation of AgtropA_{g}^{\mathrm{trop}} as a moduli space of principally polarized tropical abelian varieties, see [BMV11]. A comparison theorem, as in [CGP21, Theorem 5.8], implies the following.

Proposition 2.1.

[BBC+24, Theorem 3.1], [OO21, Corollary 2.9] For each k0k\geq 0, there is a canonical isomorphism

Hck(Agtrop)W0Hck(𝒜g).H^{k}_{c}(A_{g}^{\mathrm{trop}})\cong W_{0}H^{k}_{c}(\mathcal{A}_{g}).
Lemma 2.2.

For each g>0g>0, we have a map Ag1tropAgtropA_{g-1}^{\mathrm{trop}}\to A_{g}^{\mathrm{trop}} which is a homeomorphism onto the closed subspace of AgtropA_{g}^{\mathrm{trop}} whose complement is Pg/GLg()P_{g}/\mathrm{GL}_{g}(\mathbb{Z}).

Proof.

Consider the linear injection Sym2((g1))Sym2((g))\operatorname{Sym}^{2}((\mathbb{R}^{g-1})^{\vee})\to\operatorname{Sym}^{2}((\mathbb{R}^{g})^{\vee}) induced by the map (g1)(g)(\mathbb{R}^{g-1})^{\vee}\to(\mathbb{R}^{g})^{\vee} that extends by zero on the last basis vector in the standard ordered basis of g\mathbb{R}^{g}. It restricts to a map

(5) Pg1rtPgrtP_{g-1}^{\mathrm{rt}}\to P_{g}^{\mathrm{rt}}

which we claim is continuous. Indeed, since the Satake topology identifies PgrtP_{g}^{\mathrm{rt}} with the colimit of the cones in any admissible decomposition, it suffices to observe that the image of each perfect cone in Pg1rtP_{g-1}^{\mathrm{rt}} is a perfect cone in PgrtP_{g}^{\mathrm{rt}}. The map (5) descends to a closed embedding

Ag1tropAgtrop;A_{g-1}^{\mathrm{trop}}\hookrightarrow A_{g}^{\mathrm{trop}};

the complement of the image is Pg/GLg()P_{g}/\mathrm{GL}_{g}(\mathbb{Z}) [BBC+24, Lemma 4.9 and Proposition 4.11]. ∎

2.1.1. The tropical spectral sequence

Let ET{}^{T}\!E denote the spectral sequence on Borel–Moore homology, with rational coefficients, associated to the sequence of spaces

A0tropA1trop.\emptyset\subset A_{0}^{\mathrm{trop}}\subset A_{1}^{\mathrm{trop}}\subset\cdots.

We henceforth call ET{}^{T}\!E the tropical spectral sequence. Since AstropAs1tropA_{s}^{\mathrm{trop}}\setminus A_{s-1}^{\mathrm{trop}} is homeomorphic to Ps/GLs()P_{s}/\mathrm{GL}_{s}(\mathbb{Z}) (Lemma 2.2), we have

(6) Es,t1T=Hs+tBM(Ps/GLs();).{}^{T}\!E^{1}_{s,t}=H_{s+t}^{\mathrm{BM}}(P_{s}/\mathrm{GL}_{s}(\mathbb{Z});\mathbb{Q}).
Remark 2.3.

Alternatively, let Agtrop{}A_{g}^{\mathrm{trop}}\cup\{\infty\} denote the one-point compactification of AgtropA_{g}^{\mathrm{trop}}. Then ET{}^{T}\!E is the first quadrant spectral sequence on reduced rational homology of the sequence of pointed spaces

(7) {}A0trop{}A1trop{},\{\infty\}\subset A_{0}^{\mathrm{trop}}\cup\{\infty\}\subset A_{1}^{\mathrm{trop}}\cup\{\infty\}\cup\cdots,

with Es,t1T=Hs+t(Astrop{},As1trop{};)H~s+t(Ps/GLs(){};){}^{T}\!E^{1}_{s,t}=H_{s+t}(A_{s}^{\mathrm{trop}}\cup\{\infty\},A_{s-1}^{\mathrm{trop}}\cup\{\infty\};\mathbb{Q})\cong\widetilde{H}_{s+t}(P_{s}/\mathrm{GL}_{s}(\mathbb{Z})\cup\{\infty\};\mathbb{Q}) for s0s\geq 0. The one-point compactification point of view will play a role in Section 2.4, where the associated graded spaces of the filtered space Atrop{}A_{\infty}^{\mathrm{trop}}\cup\{\infty\} arising as the colimit of (7) shall be related to the associated graded spaces of a filtration of the space BK()BK(\mathbb{Z}).

By Poincaré duality, Es,t1T{}^{T}\!E^{1}_{s,t} is isomorphic to

(8) H(s+12)st(Ps/GLs();or).H^{\binom{s+1}{2}-s-t}(P_{s}/\mathrm{GL}_{s}(\mathbb{Z});\mathbb{Q}_{\mathrm{or}}).

Here and throughout, or\mathbb{Z}_{\mathrm{or}} is the GLs()\mathrm{GL}_{s}(\mathbb{Z})-module induced by the action on orientations of the symmetric space PsP_{s} for GLs()\mathrm{GL}_{s}(\mathbb{R}), and or:=or\mathbb{Q}_{\mathrm{or}}:=\mathbb{Z}_{\mathrm{or}}\otimes\mathbb{Q}. Note that or\mathbb{Q}_{\mathrm{or}} is nontrivial if and only if ss is even [EVGS13, Lemma 7.2]. Next, (8) is identified with group cohomology

H(s+12)st(GLs(),or).H^{\binom{s+1}{2}-s-t}(\mathrm{GL}_{s}(\mathbb{Z}),\mathbb{Q}_{\mathrm{or}}).

Since GLs()\mathrm{GL}_{s}(\mathbb{Z}) is a virtual duality group of virtual cohomological dimension (s2)\binom{s}{2}, with dualizing module Stsor,\mathrm{St}_{s}\otimes\mathbb{Z}_{\mathrm{or}}, it follows that Es,t1T{}^{T}\!E^{1}_{s,t} is isomorphic to

(9) Ht(GLs();Sts).H_{t}(\mathrm{GL}_{s}(\mathbb{Z});\mathrm{St}_{s}\otimes\mathbb{Q}).

The following proposition is a consequence of acyclicity of inflation [BBC+24, §5], as we now explain.

Proposition 2.4.

The spectral sequence

(10) Es,t1T=Ht(GLs(),Sts){}^{T}\!E^{1}_{s,t}=H_{t}(\mathrm{GL}_{s}(\mathbb{Z}),\mathrm{St}_{s}\otimes\mathbb{Q})

has exact E1E^{1}-page and hence Es,t2=0E^{2}_{s,t}=0 for all s,ts,t.

Proof.

Let

Atrop=g0Agtrop=limgAgtrop.A_{\infty}^{\mathrm{trop}}=\cup_{g\geq 0}A_{g}^{\mathrm{trop}}=\varinjlim_{g}A_{g}^{\mathrm{trop}}.

Let CBM(X)C^{\mathrm{BM}}_{*}(X) denote the locally finite chain complex associated to a space XX.

Note that ET{}^{T}\!E is the spectral sequence associated to the filtration on C=limgCBM(Agtrop)C=\varinjlim_{g}C^{\mathrm{BM}}_{*}(A_{g}^{\mathrm{trop}}) by

FsC=CBM(Astrop).F_{s}C=C^{\mathrm{BM}}_{*}(A_{s}^{\mathrm{trop}}).

For each g0g\geq 0, let I~gPgrt\widetilde{I}_{g}\subset P_{g}^{\mathrm{rt}} denote the union of the cones in the GLg()\mathrm{GL}_{g}(\mathbb{Z})-orbit of the set

(11) {σ+0egegTσΣg1perf}\{\sigma+\mathbb{R}_{\geq 0}e_{g}e_{g}^{T}\mid\sigma\in\Sigma^{\mathrm{perf}}_{g-1}\}

Let Ig=I~g/GLg()AgtropI_{g}=\widetilde{I}_{g}/\mathrm{GL}_{g}(\mathbb{Z})\subset A_{g}^{\mathrm{trop}}, henceforth referred to as the inflation locus.

Lemma 2.5.

[BBC+24, Theorem 5.15] The link LIgLI_{g} has the rational homology of a point.

Proof.

The cellular chain complex for LIgLI_{g} is the one denoted I(g)I^{(g)}_{*} in op. cit., where it is shown that I(g)I^{(g)}_{*} is acyclic.∎

Remark 2.6.

In fact LIgLI_{g} is contractible. This is not in the current literature, but can be proved using a slight modification of Yun’s Morse-theoretic enhancement of the proof of acyclicity in [BBC+24, Theorem 5.15]. See [Yun22]. Yun’s theorem deals not with LIgLI_{g} but with the link of a similarly defined matroidal coloop locus, consisting of the GLg()\mathrm{GL}_{g}(\mathbb{Z})-orbits of the unions of polyhedral cones (11) in which σ\sigma is of the form

0v1v1T,,vnvnT,\mathbb{R}_{\geq 0}\langle v_{1}v_{1}^{T},\ldots,v_{n}v_{n}^{T}\rangle,

where v1,,vnv_{1},\ldots,v_{n} are the column vectors of a g×ng\times n totally unimodular matrix.

Corollary 2.7.

HBM(Ig)=0H_{*}^{\mathrm{BM}}(I_{g})=0.

Proof.

We have HBM(Ig)=H(Ig+,)=H~(S(LIg))=H~1(LIg)H_{*}^{\mathrm{BM}}(I_{g})=H_{*}(I_{g}^{+},\infty)=\widetilde{H}_{*}(S(LI_{g}))=\widetilde{H}_{*-1}(LI_{g}), where S()S(-) denotes suspension. And H~1(LIg)=0\widetilde{H}_{*-1}(LI_{g})=0 by Lemma 2.5.∎

Now the inclusions

Ag1tropIgAgtrop,A_{g-1}^{\mathrm{trop}}\subset I_{g}\subset A_{g}^{\mathrm{trop}},

are closed and hence proper, and

HkBM(Ig)=0for all k0.H_{k}^{\mathrm{BM}}(I_{g})=0\qquad\text{for all }k\geq 0.

It follows that for each kk and for each gg,

(12) HkBM(Ag1trop)HkBM(Agtrop)H_{k}^{\mathrm{BM}}(A_{g-1}^{\mathrm{trop}})\to H_{k}^{\mathrm{BM}}(A_{g}^{\mathrm{trop}})

is the zero map, since it factors through HkBM(Ig)=0H_{k}^{\mathrm{BM}}(I_{g})=0. Now the fact that E,2T=0{}^{T}\!E^{2}_{*,*}=0 follows from the following general proposition, concluding the proof of Proposition 2.4. ∎

Proposition 2.8.

Let (C,d)(C,d) be a chain complex, and let

0=F1CF0CF1C0=F_{-1}C\subset F_{0}C\subset F_{1}C\subset\cdots

be an increasing filtration on CC, such that for each i,k0i,k\geq 0, the map

(13) Hk(FiC)Hk(Fi+1C)H_{k}(F_{i}C)\to H_{k}(F_{i+1}C)

is zero. Then the spectral sequence of the filtered chain complex

Es,t1=Hs+t(FsC,Fs1C)E^{1}_{s,t}=H_{s+t}(F_{s}C,F_{s-1}C)

collapses at Es,t2=0E^{2}_{s,t}=0 for all ss and tt.

Proof.

For each ii, the short exact sequence of chain complexes

0FiCFi+1CFi+1C/FiC00\to F_{i}C\to F_{i+1}C\to F_{i+1}C/F_{i}C\to 0

has associated long exact sequence

\displaystyle\cdots \displaystyle\longrightarrow Hk(FiC)0Hk(Fi+1C)Hk(Fi+1C,FiC)\displaystyle H_{k}(F_{i}C)\xrightarrow{0}H_{k}(F_{i+1}C)\to H_{k}(F_{i+1}C,F_{i}C)
\displaystyle\longrightarrow Hk1(FiC)0Hk1(Fi+1C)Hk1(Fi+1C,FiC)\displaystyle H_{k-1}(F_{i}C)\xrightarrow{0}H_{k-1}(F_{i+1}C)\to H_{k-1}(F_{i+1}C,F_{i}C)\to\cdots

which, by (13), splits into short exact sequences

0Hk(Fi+1C)Hk(Fi+1C,FiC)Hk1(FiC)00\to H_{k}(F_{i+1}C)\to H_{k}(F_{i+1}C,F_{i}C)\to H_{k-1}(F_{i}C)\to 0

for each k0k\geq 0. Now the rows of E1T{}^{T}\!E^{1} of the spectral sequence read

Hs+t1(Fs1C,Fs2C)Hs+t(FsC,Fs1C)Hs+t+1(Fs+1C,FsC)\cdots\leftarrow H_{s+t-1}(F_{s-1}C,F_{s-2}C)\leftarrow H_{s+t}(F_{s}C,F_{s-1}C)\leftarrow H_{s+t+1}(F_{s+1}C,F_{s}C)\leftarrow\cdots

and fit into the commuting diagram below,

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hs+t(FsC)\textstyle{H_{s+t}(F_{s}C)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hs+t1(Fs1C,Fs2C)\textstyle{H_{s+t-1}(F_{s-1}C,F_{s-2}C)}Hs+t(FsC,Fs1C)\textstyle{H_{s+t}(F_{s}C,F_{s-1}C)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hs+t+1(Fs+1C,FsC)\textstyle{H_{s+t+1}(F_{s+1}C,F_{s}C)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hs+t1(Fs1C)\textstyle{H_{s+t-1}(F_{s-1}C)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

from which it follows that the rows of E1T{}^{T}\!E^{1} are exact. Hence E2T=0{}^{T}\!E^{2}=0. ∎

Let (Cs,d)(C_{\leq s}^{*},d) denote the truncation of a cochain complex (C,d)(C^{*},d), where Csi=CiC_{\leq s}^{i}=C^{i} for isi\leq s and is zero for i>si>s. The category of cochain complexes over a field kk is equivalent to the category of graded k[x]/(x2)k[x]/(x^{2})-modules, with deg(x)=+1\deg(x)=+1. We record the following standard fact.

Lemma 2.9.

Suppose (C,d)(C^{*},d) is an acyclic cochain complex over a field kk. Then

(C,d)(sHs((Cs,d)))k[x]/(x2)(C^{*},d)\cong\Big{(}\bigoplus_{s}H^{s}((C_{\leq s}^{*},d))\Big{)}\otimes k[x]/(x^{2})

as graded k[x]/(x2)k[x]/(x^{2})-modules.

It will be convenient to consider the cohomological spectral sequence Er,T{}^{T}\!E_{r}^{*,*} dual to the tropical spectral sequence E,rT{}^{T}\!E^{r}_{*,*} in (6). The following is a consequence of Proposition 2.4, by applying Lemma 2.9 to E1T{}^{T}\!E_{1}.

Theorem 2.10.

We have an isomorphism

(E1T,d)(s0W0Hcs+t(𝒜s;))[x]/(x2)\big{(}{}^{T}\!E_{1},d\big{)}\cong\Big{(}\bigoplus_{s\geq 0}W_{0}H^{s+t}_{c}(\mathcal{A}_{s};\mathbb{Q})\Big{)}\otimes\mathbb{Q}[x]/(x^{2})

of 2\mathbb{Z}^{2}-graded cochain complexes with differential in degree (1,0)(1,0). Here, W0Hcs+t(𝒜s;)W_{0}H^{s+t}_{c}(\mathcal{A}_{s};\mathbb{Q}) is in bidegree (s,t)(s,t), and xx is in bidegree (1,0)(1,0).

Proof.

Proposition 2.4 shows that E1T{}^{T}\!E_{1} is acyclic. The truncation of E1T{}^{T}\!E_{1} after the first ss columns is the E1E_{1}-page of the spectral sequence that converges on E2E_{2} to W0Hcs+t(𝒜s;)W_{0}H^{s+t}_{c}(\mathcal{A}_{s};\mathbb{Q}), supported in column ss. The statement then follows from Lemma 2.9. ∎

In Proposition 4.4 we shall upgrade Theorem 2.10 to an isomorphism of graded-commutative algebras over \mathbb{Q}.

2.2. The Waldhausen construction of KK-theory

We briefly review here the Waldhausen SS_{\bullet}-construction of KK-theory, following the survey [Wei13, IV.8]. The construction applies to arbitrary Waldhausen categories. We shall first recall the definition of a Waldhausen category. We do this only briefly, keeping in mind that our main example shall be the category of finitely generated projective left modules over a ring, as in Example 2.11 below. We refer to [Wei13, II.9] for full and precise definitions.

A category with cofibrations is a category 𝒞\mathcal{C} equipped with a special subcategory of morphisms co(𝒞)\mathrm{co}(\mathcal{C}) called cofibrations. Every isomorphism is a cofibration. There is a distinguished 0 object in 𝒞\mathcal{C} and the unique map from 0 to any object is a cofibration. Cofibrations are preserved under pushouts along arbitrary morphisms; in particular, such pushouts exist. We denote cofibrations with a feathered arrow \rightarrowtail.

A Waldhausen category is a category with cofibrations together with a family w(𝒞)w(\mathcal{C}) of morphisms, called weak equivalences. Weak equivalences will be denoted \xrightarrow{\sim}. All isomorphisms are weak equivalences. Weak equivalences are closed under composition and satisfy the following gluing condition: for every commutative diagram

C\textstyle{C\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\sim}A\textstyle{A\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\sim}B\textstyle{B\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\sim}C\textstyle{C^{\prime}}A\textstyle{A^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}B\textstyle{B^{\prime}}

the induced map of pushouts

BACBACB\cup_{A}C\to B^{\prime}\cup_{A^{\prime}}C^{\prime}

is a weak equivalence. The following is our main example of interest.

Example 2.11.

Let RR be a ring; we do not require that RR is commutative. Let ProjR\mathrm{Proj}_{R} denote the category of finitely generated projective left RR-modules, let co(ProjR)\mathrm{co}(\mathrm{Proj}_{R}) be the injections with projective cokernel, and let w(ProjR)w(\mathrm{Proj}_{R}) be the isomorphisms. Then (ProjR,co(ProjR),w(ProjR))(\mathrm{Proj}_{R},\mathrm{co}(\mathrm{Proj}_{R}),w(\mathrm{Proj}_{R})) is a Waldhausen category.

We now review Waldhausen’s SS_{\bullet}-construction of KK-theory, following [Wei13, IV.8]. Let 𝒞\mathcal{C} be any Waldhausen category. For each p0p\geq 0, define a category Sp(𝒞)S_{p}(\mathcal{C}) whose objects are commutative diagrams

(14) 0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}P0,1\textstyle{P_{0,1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}P0,2\textstyle{P_{0,2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}P0,p\textstyle{P_{0,p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}P1,2\textstyle{P_{1,2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}P1,p\textstyle{P_{1,p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\vdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Pp1,p\textstyle{P_{p-1,p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

in which the horizontal morphisms are cofibrations and Pi,jP_{i,j} is a chosen cokernel for the horizontal map PiPjP_{i}\to P_{j}, where Pi:=P0,iP_{i}:=P_{0,i}. The arrows in Sp(𝒞)S_{p}(\mathcal{C}) are morphisms PQP_{\bullet}\to Q_{\bullet} of such diagrams that are weak equivalences. By a weak equivalence, we mean a morphism PQP_{\bullet}\to Q_{\bullet} such that each component Pi,jQi,jP_{i,j}\to Q_{i,j} is a weak equivalence.

Remark 2.12.

In other sources such as [Wei13, IV.8], what we write as Sp(𝒞)S_{p}(\mathcal{C}) above is written as wSp(𝒞)wS_{p}(\mathcal{C}), while Sp(𝒞)S_{p}(\mathcal{C}) is reserved for the category whose objects are diagrams  (14) and whose morphisms are morphisms of diagrams, with no restrictions. This more general notion of morphism in Sp(𝒞)S_{p}(\mathcal{C}) is useful when iterating the SS_{\bullet}-construction, which we will not need here. Since the only morphisms in Sp(𝒞)S_{p}(\mathcal{C}) relevant to us are the weak equivalences, we shall write Sp(𝒞)S_{p}(\mathcal{C}) instead of wSp(𝒞)wS_{p}(\mathcal{C}) for the category whose objects are triangular diagrams (14) and whose morphisms are weak equivalences.

Returning to Waldhausen’s SS_{\bullet}-construction, we define functors di:S(𝒞)S1(𝒞)d_{i}\colon S_{\bullet}(\mathcal{C})\to S_{\bullet-1}(\mathcal{C}) for each ii by deleting both the iith row and the iith column in (14) (i.e., the row with objects labeled Pi,jP_{i,j} and the column with objects labeled Pk,iP_{k,i}) and then re-indexing the remaining terms. Similarly, we define functors si:S(𝒞)S+1(𝒞)s_{i}\colon S_{\bullet}(\mathcal{C})\to S_{\bullet+1}(\mathcal{C}) by duplicating both the iith row and the iith column in (14), inserting identity morphisms, and then re-indexing. In this way, S(𝒞)S_{\bullet}(\mathcal{C}) is a simplicial object in the category of categories. Taking the nerve yields a bisimplicial set

NS(𝒞):Δop×Δop\displaystyle N_{\bullet}S_{\bullet}(\mathcal{C})\colon\Delta^{\mathrm{op}}\times\Delta^{\mathrm{op}} 𝖲𝖾𝗍\displaystyle\to\mathsf{Set}
([p],[q])\displaystyle([p],[q]) NqSp(𝒞).\displaystyle\mapsto N_{q}S_{p}(\mathcal{C}).
Remark 2.13.

In order to get a bisimplicial set, the category 𝒞\mathcal{C} must be small, i.e. there is a set of objects. In that case S(𝒞)S_{\bullet}(\mathcal{C}) is also small. The main examples, such as finitely generated projective left modules over some ring, are essentially small but not small if one takes all such objects. In those cases one chooses (often tacitly) a small category 𝒞small\mathcal{C}^{\mathrm{small}} equivalent to 𝒞\mathcal{C} and applies the Waldhausen construction to that. The resulting space |NS(𝒞small)||N_{\bullet}S_{\bullet}(\mathcal{C}^{\mathrm{small}})| is independent of this choice, up to homotopy equivalence.

In the case 𝒞=Proj\mathcal{C}=\mathrm{Proj}_{\mathbb{Z}} of finitely generated projective modules over \mathbb{Z}, we may make the following explicit choice of (Proj)small(\mathrm{Proj}_{\mathbb{Z}})^{\mathrm{small}}. The set of objects of (Proj)small(\mathrm{Proj}_{\mathbb{Z}})^{\mathrm{small}} is ={0,1,2,}\mathbb{N}=\{0,1,2,\ldots\}. The set Mor(Proj)small(a,b)\mathrm{Mor}_{(\mathrm{Proj}_{\mathbb{Z}})^{\mathrm{small}}}(a,b) is the set of b×ab\times a integer matrices. Composition is given by matrix multiplication. This category may be further equipped with a symmetric monoidal structure corresponding to direct sum of \mathbb{Z}-modules: the product is given on objects by m(a,b)=a+bm(a,b)=a+b, and on morphisms by block sum of matrices. The associator

(Proj)small×(Proj)small×(Proj)small{(\mathrm{Proj}_{\mathbb{Z}})^{\mathrm{small}}\times(\mathrm{Proj}_{\mathbb{Z}})^{\mathrm{small}}\times(\mathrm{Proj}_{\mathbb{Z}})^{\mathrm{small}}}(Proj)small{(\mathrm{Proj}_{\mathbb{Z}})^{\mathrm{small}}}

is given on (a,b,c)(a,b,c) by the identity matrix of size a+b+ca+b+c. The symmetry natural isomorphism is given on (a,b)(Proj)small×(Proj)small(a,b)\in(\mathrm{Proj}_{\mathbb{Z}})^{\mathrm{small}}\times(\mathrm{Proj}_{\mathbb{Z}})^{\mathrm{small}} by the (a+b)×(a+b)(a+b)\times(a+b) block matrix

(0IdaIdb0).\left(\begin{matrix}0&\mathrm{Id}_{a}\\ \mathrm{Id}_{b}&0\end{matrix}\right).

In this case, sending nnn\mapsto\mathbb{Z}^{n} extends to a symmetric monoidal equivalence from (Proj)small(\mathrm{Proj}_{\mathbb{Z}})^{\mathrm{small}} to the category Proj\mathrm{Proj}_{\mathbb{Z}} of all finitely generated free \mathbb{Z}-modules (equipped with the cartesian symmetric monoidal structure, i.e., direct sum).

2.3. The Quillen spectral sequence

Recall the category ProjR\mathrm{Proj}_{R} of finitely generated projective left RR-modules over a ring RR, considered as a Waldhausen category as in Example 2.11. Then

(15) BK(R):=|NS(ProjR)|, and Ki(R):=πi+1BK(R).BK(R):=|N_{\bullet}S_{\bullet}(\mathrm{Proj}_{R})|,\mbox{ \ \ and \ \ }K_{i}(R):=\pi_{i+1}BK(R).

By (15) and its degree shift, BK(R)BK(R) is a de-looping of the KK-theory space of RR.

We now take R=R=\mathbb{Z}. Consider the following filtrations on the objects of Proj\mathrm{Proj}_{\mathbb{Z}} and S(Proj)S_{\bullet}(\mathrm{Proj}_{\mathbb{Z}}). First, an object PP is in FnF_{n} if rank(P)n\mathrm{rank}(P)\leq n. This filtration induces a filtration on the objects of Sp(Proj)S_{p}(\mathrm{Proj}_{\mathbb{Z}}), in which an object of Sp(Proj)S_{p}(\mathrm{Proj}_{\mathbb{Z}}), which is a triangular diagram (14), is in FnF_{n} if the top right projective module is in FnF_{n}, i.e., has rank at most nn. Since that rank is non-increasing under face and degeneracy maps, and is preserved under weak equivalences, it follows that the filtration on objects of Sp(Proj)S_{p}(\mathrm{Proj}_{\mathbb{Z}}), for each p0p\geq 0 yields a filtered bisimplicial set NS(Proj)N_{\bullet}S_{\bullet}(\mathrm{Proj}_{\mathbb{Z}}). Hence, we obtain an exhaustive filtration of based spaces

(16) F0BK()FnBK()Fn+1BK()BK(),F_{0}BK(\mathbb{Z})\subset\cdots\subset F_{n}BK(\mathbb{Z})\subset F_{n+1}BK(\mathbb{Z})\subset\cdots\subset BK(\mathbb{Z}),

where F0BK()F_{0}BK(\mathbb{Z}) is contractible.

The functor Proj×ProjProj\mathrm{Proj}_{\mathbb{Z}}\times\mathrm{Proj}_{\mathbb{Z}}\to\mathrm{Proj}_{\mathbb{Z}} defined by direct sum of \mathbb{Z}-modules induces a map

BK()×BK()𝑚BK()BK(\mathbb{Z})\times BK(\mathbb{Z})\xrightarrow{m}BK(\mathbb{Z})

which is filtered in the sense that m(Fa(BK())×Fb(BK()))Fa+bBK()m(F_{a}(BK(\mathbb{Z}))\times F_{b}(BK(\mathbb{Z})))\subset F_{a+b}BK(\mathbb{Z}), and hence induces a product on the spectral sequence associated to the filtration (see Proposition 3.13 below for more details).

Definition 2.14.

Let E,Q{}^{Q}\!E^{*}_{*,*} denote the homology spectral sequence associated to the rank filtration on BK()BK(\mathbb{Z}), called the Quillen spectral sequence.

A filtration of BK()BK(\mathbb{Z}) equivalent to the one above, and more generally a filtration of BK(𝒪F)BK(\mathcal{O}_{F}) for a number field FF, was used by Quillen in [Qui73, pp. 179–199] for his proof that Ki(𝒪F)K_{i}(\mathcal{O}_{F}) is a finitely generated abelian group for all ii\in\mathbb{N}. Quillen’s construction of the filtration used the “QQ-construction” model for BK(𝒪F)BK(\mathcal{O}_{F}), and he also identified the associated graded with homology of the Steinberg module.

Proposition 2.15.

There is a canonical isomorphism

(17) Es,t1QHt(GLs();Sts)H(BK()).{}^{Q}\!E^{1}_{s,t}\cong H_{t}(\mathrm{GL}_{s}(\mathbb{Z});\mathrm{St}_{s}\otimes\mathbb{Q})\Rightarrow H_{\ast}(BK(\mathbb{Z})).

With respect to this isomorphism, the product on the E1E^{1} page is induced by the block sum homomorphism GLs()×GLs()GLs+s()\mathrm{GL}_{s}(\mathbb{Z})\times\mathrm{GL}_{s^{\prime}}(\mathbb{Z})\to\mathrm{GL}_{s+s^{\prime}}(\mathbb{Z}) and the GLs()×GLs()\mathrm{GL}_{s}(\mathbb{Z})\times\mathrm{GL}_{s^{\prime}}(\mathbb{Z})-equivariant map StsStsSts+s\mathrm{St}_{s}\otimes\mathrm{St}_{s^{\prime}}\to\mathrm{St}_{s+s^{\prime}} induced by the “block sum” map of spaces Ts()Ts()Ts+s()T_{s}(\mathbb{Q})\ast T_{s^{\prime}}(\mathbb{Q})\to T_{s+s^{\prime}}(\mathbb{Q}).

The Tits building Ts+s()T_{s+s^{\prime}}(\mathbb{Q}) is the (nerve of the) poset of non-zero proper linear subspaces of s+s=ss\mathbb{Q}^{s+s^{\prime}}=\mathbb{Q}^{s}\oplus\mathbb{Q}^{s^{\prime}}. Identifying s\mathbb{Q}^{s} with s{0}s+s\mathbb{Q}^{s}\oplus\{0\}\subset\mathbb{Q}^{s+s^{\prime}} and s\mathbb{Q}^{s^{\prime}} with {0}ss+s\{0\}\oplus\mathbb{Q}^{s^{\prime}}\subset\mathbb{Q}^{s+s^{\prime}} leads to a canonical embedding of simplicial complexes

Ts()Ts()\displaystyle T_{s}(\mathbb{Q})\ast T_{s^{\prime}}(\mathbb{Q}) Ts+s()\displaystyle\hookrightarrow T_{s+s^{\prime}}(\mathbb{Q})
(VTs())\displaystyle(V\in T_{s}(\mathbb{Q})) V{0}\displaystyle\mapsto V\oplus\{0\}
(VTs())\displaystyle(V^{\prime}\in T_{s^{\prime}}(\mathbb{Q})) {0}V,\displaystyle\mapsto\{0\}\oplus V^{\prime},

and this is the “block sum” map referred to in the proposition. It is evidently equivariant with respect to the usual block sum operation of matrices

(A,B)(A00B).(A,B)\mapsto\begin{pmatrix}A&0\\ 0&B\end{pmatrix}.

We remark that for s=0s=0, one should take St0=\mathrm{St}_{0}=\mathbb{Q} in the above formula for the E1E^{1}-page, and that the canonical isomorphism =H0(GL0();St0)\mathbb{Q}=H_{0}(\mathrm{GL}_{0}(\mathbb{Z});\mathrm{St}_{0}) gives a two-sided unit for the product.

Proof.

As mentioned above this is essentially due to Quillen, although his construction of the spectral sequence and identification of the E1E^{1}-page with the homology of the Steinberg module used a different model of BK()BK(\mathbb{Z}). His model, the QQ-construction, compares to the Waldhausen construction by an explicit homotopy equivalence explained in [Wal85, Section 1.9]. One may verify that this homotopy equivalence induces a homotopy equivalence in each filtration degree.

In Section 2.4, we need to use what the isomorphism Es,t1QHt(GLs();Sts){}^{Q}\!E^{1}_{s,t}\cong H_{t}(\mathrm{GL}_{s}(\mathbb{Z});\mathrm{St}_{s}\otimes\mathbb{Q}) is, so we sketch a direct construction which makes no mention of the QQ-construction. Let us write

Spn(Proj)=FnSp(Proj)Fn1Sp(Proj)S^{n}_{p}(\mathrm{Proj}_{\mathbb{Z}})=F_{n}S_{p}(\mathrm{Proj}_{\mathbb{Z}})\setminus F_{n-1}S_{p}(\mathrm{Proj}_{\mathbb{Z}})

for the full subgroupoid on those triangular diagrams (14) in which P0,pnP_{0,p}\cong\mathbb{Z}^{n}. Then the filtration quotients are described, in bidegree (p,q)(p,q), by the bijection of pointed sets

(18) FnNqSp(Proj)Fn1NqSp(Proj)NqSpn(Proj){}.\frac{F_{n}N_{q}S_{p}(\mathrm{Proj}_{\mathbb{Z}})}{F_{n-1}N_{q}S_{p}(\mathrm{Proj}_{\mathbb{Z}})}\cong N_{q}S^{n}_{p}(\mathrm{Proj}_{\mathbb{Z}})\cup\{\infty\}.

Here, the quotient S/TS/T of a set SS by a subset TT, or more generally along a morphism TST\to S, is understood as the pushout of the diagram {}TS\{\ast\}\leftarrow T\rightarrow S. This pushout S/TS/T is naturally regarded as a pointed set with distinguished element corresponding to the image of {}S/T\{\ast\}\to S/T. On the right hand side, \infty denotes a disjoint basepoint corresponding to the collapsed subset. There are well defined functors di:Spn(Proj)Sp1n(Proj)d_{i}\colon S_{p}^{n}(\mathrm{Proj}_{\mathbb{Z}})\to S^{n}_{p-1}(\mathrm{Proj}_{\mathbb{Z}}) for 0<i<p0<i<p, defined by deleting the iith row and column of the triangular diagram, but the outer faces d0d_{0} and dpd_{p} will not always land in the subgroupoid Sp1n(Proj)FnSp(Proj)S^{n}_{p-1}(\mathrm{Proj}_{\mathbb{Z}})\subset F_{n}S_{p}(\mathrm{Proj}_{\mathbb{Z}}) because the rank may drop (which happens unless P0,10P_{0,1}\cong 0 or Pp1,p0P_{p-1,p}\cong 0, respectively). In the bisimplicial set (18), the face operators in the pp-direction will send a simplex to the basepoint \infty when this happens. We deduce that En,1E^{1}_{n,*} is calculated as the homology of the total complex associated to the double complex with

Cp,q=NqSpn(Proj)=NqSpn(Proj){}{},C_{p,q}=\mathbb{Q}\langle N_{q}S^{n}_{p}(\mathrm{Proj}_{\mathbb{Z}})\rangle=\frac{\mathbb{Q}\langle N_{q}S^{n}_{p}(\mathrm{Proj}_{\mathbb{Z}})\cup\{\infty\}\rangle}{\mathbb{Q}\langle\{\infty\}\rangle},

where X\mathbb{Q}\langle X\rangle denotes the free \mathbb{Q}-module on a set XX, and boundary maps defined as alternating sum of face maps as usual.

The groupoid Sp(Proj)S_{p}(\mathrm{Proj}_{\mathbb{Z}}) comes with a functor FF to the groupoid of finitely generated projective \mathbb{Z}-modules and their isomorphisms, defined by sending a triangular diagram (14) to P0,pP_{0,p}. The categorical fiber (Fn)(F\downarrow\mathbb{Z}^{n}) over the object n\mathbb{Z}^{n} is the groupoid whose objects are triangular diagrams equipped with a specified isomorphism P0,pnP_{0,p}\to\mathbb{Z}^{n}, and where morphisms from one triangular diagram to another is an isomorphism of diagrams compatible with the two specified isomorphisms from upper right entries to n\mathbb{Z}^{n}. If we denote this groupoid S~pn(Proj)\widetilde{S}^{n}_{p}(\mathrm{Proj}_{\mathbb{Z}}), then there is a forgetful map

(19) NqS~pn(Proj)NqSpn(Proj)N_{q}\widetilde{S}^{n}_{p}(\mathrm{Proj}_{\mathbb{Z}})\to N_{q}S^{n}_{p}(\mathrm{Proj}_{\mathbb{Z}})

which identifies NqSpn(Proj)N_{q}S^{n}_{p}(\mathrm{Proj}_{\mathbb{Z}}) with the set of orbits for the free GLn()\mathrm{GL}_{n}(\mathbb{Z})-action on the set NqS~pn(Proj)N_{q}\widetilde{S}^{n}_{p}(\mathrm{Proj}_{\mathbb{Z}}), defined by acting by post-composition on the specified isomorphisms P0,pnP_{0,p}\cong\mathbb{Z}^{n}. There is also a map of sets

(20) NqS~pn(Proj)π0(NS~pn(Proj))N_{q}\widetilde{S}^{n}_{p}(\mathrm{Proj}_{\mathbb{Z}})\to\pi_{0}(N_{\bullet}\widetilde{S}^{n}_{p}(\mathrm{Proj}_{\mathbb{Z}}))

sending a qq-simplex to the path component containing it. Regarding pp as fixed, the maps (20) assemble to a map of simplicial sets from [q]NqS~pn(Proj)[q]\mapsto N_{q}\widetilde{S}^{n}_{p}(\mathrm{Proj}_{\mathbb{Z}}) to the constant simplicial set [q]π0(NS~pn(Proj))[q]\mapsto\pi_{0}(N_{\bullet}\widetilde{S}^{n}_{p}(\mathrm{Proj}_{\mathbb{Z}})). This map is a weak homotopy equivalence because automorphisms in the groupoid S~pn(Proj)\widetilde{S}^{n}_{p}(\mathrm{Proj}_{\mathbb{Z}}) are uniquely determined by their action on the top-right module P0,p=nP_{0,p}=\mathbb{Z}^{n} and hence all automorphism groups in this groupoid are trivial.

The set π0(NS~pn(Proj))\pi_{0}(N_{\bullet}\widetilde{S}^{n}_{p}(\mathrm{Proj}_{\mathbb{Z}})) is in bijection with the set of flags 0P0,1P0,p1n0\subset P_{0,1}\subset\dots\subset P_{0,p-1}\subset\mathbb{Z}^{n} of saturated submodules. Such a flag is non-degenerate as a pp-simplex of [p]π0(NS~pn(Proj))[p]\mapsto\pi_{0}(N_{\bullet}\widetilde{S}^{n}_{p}(\mathrm{Proj}_{\mathbb{Z}})) if and only if none of the inclusions are equalities. For p2p\geq 2, this set of non-degenerate pp-simplices is in bijection with the set of (p2)(p-2)-dimensional faces of the Tits building Tn()T_{n}(\mathbb{Q}). Writing

C~p,q=NqS~pn(Proj)=NqS~pn(Proj){}{},\widetilde{C}_{p,q}=\mathbb{Q}\langle N_{q}\widetilde{S}^{n}_{p}(\mathrm{Proj}_{\mathbb{Z}})\rangle=\frac{\mathbb{Q}\langle N_{q}\widetilde{S}^{n}_{p}(\mathrm{Proj}_{\mathbb{Z}})\cup\{\infty\}\rangle}{\mathbb{Q}\langle\{\infty\}\rangle},

we deduce that the maps

TotpC~,C~p2(Tn())\operatorname{Tot}_{p}\widetilde{C}_{*,*}\to\widetilde{C}_{p-2}(T_{n}(\mathbb{Q}))

that are induced by (20) on C~p,0\widetilde{C}_{p,0} and zero on C~p,pp\widetilde{C}_{p^{\prime},p-p^{\prime}} for ppp^{\prime}\neq p assemble to a quasi-isomorphism of chain complexes, and hence, for n1n\geq 1, an isomorphism

Hp(TotC~,){Stn()for p=n0otherwise.H_{p}(\operatorname{Tot}\widetilde{C}_{*,*})\cong\begin{cases}\mathrm{St}_{n}(\mathbb{Q})\otimes\mathbb{Q}&\text{for $p=n$}\\ 0&\text{otherwise}.\end{cases}

The maps (19) induce an isomorphism of double complexes

[GLn()]C~,C,\mathbb{Q}\otimes_{\mathbb{Q}[\mathrm{GL}_{n}(\mathbb{Z})]}\widetilde{C}_{*,*}\xrightarrow{\cong}C_{*,*}

and hence an isomorphism of associated total complexes. Since Tot(C~,)\operatorname{Tot}(\widetilde{C}_{*,*}) is a complex of free [GLn()]\mathbb{Q}[\mathrm{GL}_{n}(\mathbb{Z})]-modules, we can use it to compute group homology as

Hp(GLn();Stn())\displaystyle H_{p}(\mathrm{GL}_{n}(\mathbb{Z});\mathrm{St}_{n}(\mathbb{Q})) =Torp[GLn()](,Stn())=Hpn([GLn()]TotC~,)\displaystyle=\mathrm{Tor}_{p}^{\mathbb{Q}[\mathrm{GL}_{n}(\mathbb{Z})]}(\mathbb{Q},\mathrm{St}_{n}(\mathbb{Q}))=H_{p-n}(\mathbb{Q}\otimes_{\mathbb{Q}[\mathrm{GL}_{n}(\mathbb{Z})]}\operatorname{Tot}\widetilde{C}_{*,*})
=Hpn(TotC,)=Hpn(Fn(BK()),Fn1(BK());),\displaystyle=H_{p-n}(\operatorname{Tot}C_{*,*})=H_{p-n}(F_{n}(BK(\mathbb{Z})),F_{n-1}(BK(\mathbb{Z}));\mathbb{Q}),

as required. The assertion about the algebra structure follows by tracing isomorphisms, since direct sum of \mathbb{Z}-modules corresponds to block sum of matrices upon choosing an isomorphism to n\mathbb{Z}^{n} for some nn. ∎

Remark 2.16.

We have stated the theorem and proof above for the ring R=R=\mathbb{Z}, but the rank filtration evidently makes sense for any ring RR, by considering the full subcategories Fg(ProjR)F_{g}(\mathrm{Proj}_{R}) on those projective modules which arise as summands of RgR^{\oplus g}.

The formula for the E1E^{1}-page generalizes to any Dedekind domain RR, for instance R=𝒪ER=\mathcal{O}_{E} for a number field EE, or if RR is any field. If KK is the fraction field of a Dedekind domain RR, then saturated subspaces of a finitely generated projective RR-module PP of rank ss are in bijection with linear subspaces of PRKKsP\otimes_{R}K\cong K^{\oplus s}, and the Tits building T(PRK)Ts(K)T(P\otimes_{R}K)\approx T_{s}(K) is defined as the nerve of the partially ordered set of non-zero proper linear subspaces of PRKP\otimes_{R}K. The geometric realization of this partially ordered set has the homotopy type of a wedge of (s2)(s-2)-spheres by the Solomon–Tits theorem, and the Steinberg module can be defined as St(PRK)=H~s2(|T(PRK)|)Sts(K)\operatorname{St}(P\otimes_{R}K)=\widetilde{H}_{s-2}(|T(P\otimes_{R}K)|)\cong\operatorname{St}_{s}(K). The proof above goes through in this generality with only minor modifications, giving a Quillen spectral sequence of the form

Es,t1Q=PHt(GL(P);St(PRK))Hs+t(BK(R)){}^{Q}\!E^{1}_{s,t}=\bigoplus_{P}H_{t}(\mathrm{GL}(P);\operatorname{St}(P\otimes_{R}K))\Rightarrow H_{s+t}(BK(R))

where PP ranges over a set of representatives for the isomorphism classes of projective modules PP over RR of rank ss. For a completely general ring we still have a spectral sequence associated to the rank filtration, but not a useful formula for its E1E^{1}-page.

In Section 5.1, we explain a construction of KK-theory for graphs that is closely related to the SS_{\bullet}-construction for Waldhausen categories. Several variants are possible. The variant we use, which seems most closely related to the Quillen spectral sequence, does not fit the definitions of a Waldhausen category, but is similar in spirit.

2.4. Comparison of tropical and Quillen spectral sequences

Recall that the tropical spectral sequence and the Quillen spectral sequence have isomorphic E1E^{1}-pages

Es,t1THt(GLs();Sts)Es,t1Q.{}^{T}\!E^{1}_{s,t}\cong H_{t}(\mathrm{GL}_{s}(\mathbb{Z});\operatorname{St}_{s}\otimes\mathbb{Q})\cong{}^{Q}\!E^{1}_{s,t}.

The goal of this subsection is to make the implied isomorphism of E1E^{1}-pages more explicit. Recall that Borel–Moore homology of a reasonable space agrees with the relative homology of its one-point compactification, for instance HsBM(Agtrop)=Hs(Agtrop{},)H_{s}^{\mathrm{BM}}(A_{g}^{\mathrm{trop}})=H_{s}(A_{g}^{\mathrm{trop}}\cup\{\infty\},\infty). Therefore, the tropical spectral sequence can be viewed as the spectral sequence associated (in ordinary singular homology) to the filtered space

Atrop{}=colimgAgtrop{},A_{\infty}^{\mathrm{trop}}\cup\{\infty\}=\operatorname*{colim}_{g\to\infty}A_{g}^{\mathrm{trop}}\cup\{\infty\},

the direct limit of the one-point compactifications of the locally compact space AgtropA_{g}^{\mathrm{trop}}. The associated graded has

Grg(Atrop{})(Pg/GLg()){},\operatorname{Gr}_{g}(A_{\infty}^{\mathrm{trop}}\cup\{\infty\})\cong(P_{g}/\mathrm{GL}_{g}(\mathbb{Z}))\cup\{\infty\},

the one-point compactification of Pg/GLg()P_{g}/\mathrm{GL}_{g}(\mathbb{Z}).

In this subsection we will construct an explicit zig-zag of (rational) equivalences

(21) Grg(BK())|NT(g){}|Grg(Atrop{}))\operatorname{Gr}_{g}(BK(\mathbb{Z}))\xleftarrow{\simeq}|N_{\bullet}T_{\bullet}(\mathbb{Q}^{g})\cup\{\infty\}|\xrightarrow{\simeq_{\mathbb{Q}}}\operatorname{Gr}_{g}(A_{\infty}^{\mathrm{trop}}\cup\{\infty\}))

in which the middle term |NT(g){}||N_{\bullet}T_{\bullet}(\mathbb{Q}^{g})\cup\{\infty\}| will be defined in this section. The reduced homology of the rightmost space agrees with the Borel–Moore homology of Pg/GLg()P_{g}/\mathrm{GL}_{g}(\mathbb{Z}) and hence is the E1E^{1}-page of the tropical spectral sequence. The reduced homology of the leftmost space in the zig-zag is manifestly the E1E^{1}-page of the Quillen spectral sequence, and for both spectral sequences we have identified the E1E^{1}-page as Es,t1=Ht(GLs(),Sts)E^{1}_{s,t}=H_{t}(\mathrm{GL}_{s}(\mathbb{Z}),\operatorname{St}_{s}\otimes\mathbb{Q}). The purpose of discussing this zig-zag is to evaluate certain pairings: given classes

wEs,t1Q\displaystyle w\in{}^{Q}\!E^{1}_{s,t} Ht(GLs();Sts)\displaystyle\cong H_{t}(\mathrm{GL}_{s}(\mathbb{Z});\operatorname{St}_{s}\otimes\mathbb{Q})
ωE1s,tT\displaystyle\omega\in{}^{T}\!E_{1}^{s,t}\otimes\mathbb{R} Hom(Ht(GLs();Sts),)\displaystyle\cong\operatorname{Hom}(H_{t}(\mathrm{GL}_{s}(\mathbb{Z});\operatorname{St}_{s}),\mathbb{R})

we wish to make sense of the pairing w,ω\langle w,\omega\rangle\in\mathbb{R}. To do this we need to pin down the isomorphism Es,t1QEs,t1T{}^{Q}\!E^{1}_{s,t}\to{}^{T}\!E^{1}_{s,t}, which is what the zig-zag (21) is useful for. In practice, the homology class ww will be represented by a somewhat explicit cycle in Grg(BK())\operatorname{Gr}_{g}(BK(\mathbb{Z})) arising from graphs, and the cohomology class ω\omega will be given by an explicit GLg()\mathrm{GL}_{g}(\mathbb{Z})-equivariant differential form on PgP_{g}. In this situation, the strategy will be to transfer ww along the zig-zag to get an interpretation as a cycle in (Agtrop{})/(Ag1trop{})(A_{g}^{\mathrm{trop}}\cup\{\infty\})/(A_{g-1}^{\mathrm{trop}}\cup\{\infty\}) and evaluate the pairing by performing an integral.

Before constructing the zig-zag (21), let us point out that this must happen at the level of associated gradeds and cannot be lifted to the level of filtered spaces, since the tropical spectral sequence collapses to E2=0E^{2}=0 while the Quillen spectral sequence does not.

We first define the space in the middle.

Definition 2.17.

For a finite-dimensional \mathbb{Q}-vector space VV, let N0Tp(V)N_{0}T_{p}(V) be the set of pairs (ApA0,<)(A_{p}\subset\dots\subset A_{0},<) where

ApA0V{0}A_{p}\subset\dots\subset A_{0}\subset V^{\vee}\setminus\{0\}

is a flag of non-zero vectors in the dual vector space, << is a total order on A0A_{0}, such that

  1. (1)

    AiA_{i} is finite for all ii,

  2. (2)

    Ap=A_{p}=\emptyset,

  3. (3)

    span(A0)=V\mathrm{span}(A_{0})=V^{\vee}.

Let N0Tp(V){}N_{0}T_{p}(V)\cup\{\infty\} denote the pointed set obtained by adding a disjoint base point, which we denote \infty, and define maps of pointed sets

di:N0Tp(V){}N0Tp1(V){}d_{i}\colon N_{0}T_{p}(V)\cup\{\infty\}\to N_{0}T_{p-1}(V)\cup\{\infty\}

for 0ip0\leq i\leq p by the formula

di(ApA0,<)={if i=0<p and span(A1)Vif i=p>0 and Ap1(ApAi^A0,<)otherwise.d_{i}(A_{p}\subset\dots\subset A_{0},<)=\begin{cases}\infty&\text{if $i=0<p$ and $\mathrm{span}(A_{1})\neq V^{\vee}$}\\ \infty&\text{if $i=p>0$ and $A_{p-1}\neq\emptyset$}\\ (A_{p}\subset\dots\subset\widehat{A_{i}}\subset\dots\subset A_{0},<)&\text{otherwise.}\end{cases}

There are also degeneracy maps si:N0Tp1(V){}N0Tp(V){}s_{i}\colon N_{0}T_{p-1}(V)\cup\{\infty\}\to N_{0}T_{p}(V)\cup\{\infty\} defined by duplicating AiA_{i}, making [p]N0Tp(V){}[p]\mapsto N_{0}T_{p}(V)\cup\{\infty\} into a simplicial object in the category of pointed sets.

There is an evident action of GLg()\mathrm{GL}_{g}(\mathbb{Z}) on N0Tp(g)N_{0}T_{p}(\mathbb{Q}^{g}), which we can use to turn N0Tp(g)N_{0}T_{p}(\mathbb{Q}^{g}) into the objects of a groupoid Tp(g)T_{p}(\mathbb{Q}^{g}): the set of morphisms from (ApA0)(A_{p}\subset\dots\subset A_{0}) to (ApA0)(A^{\prime}_{p}\subset\dots\subset A^{\prime}_{0}) is the set of matrices XGLg()X\in\mathrm{GL}_{g}(\mathbb{Z}) with the property that X.(ApA0)=(ApA0)X.(A_{p}\subset\dots\subset A_{0})=(A^{\prime}_{p}\subset\dots\subset A^{\prime}_{0}). (This action is of course the restriction of an action of GLg()\mathrm{GL}_{g}(\mathbb{Q}), but for our purposes we need the action by integer matrices only.) Explicitly,

NqTp(g)=(GLg())q×N0Tp(g),N_{q}T_{p}(\mathbb{Q}^{g})=(\mathrm{GL}_{g}(\mathbb{Z}))^{q}\times N_{0}T_{p}(\mathbb{Q}^{g}),

with face maps in the qq-direction given by

di(X1,,Xq,(ApA0,<))={(X2,,Xq,(ApA0,<))for i=0(X1,,XiXi+1,,Xq,(ApA0,<))for 0<i<p(X1,,Xq1,Xq.(ApA0,<))\begin{aligned} d_{i}(X_{1},\dots,X_{q},(A_{p}\subset\dots\subset A_{0},<))=\begin{cases}(X_{2},\dots,X_{q},(A_{p}\subset\dots\subset A_{0},<))&\text{for $i=0$}\\ (X_{1},\dots,X_{i}X_{i+1},\dots,X_{q},(A_{p}\subset\dots\subset A_{0},<))&\text{for $0<i<p$}\\ (X_{1},\dots,X_{q-1},X_{q}.(A_{p}\subset\dots\subset A_{0},<))\end{cases}\end{aligned}

The notation NT(g){}N_{\bullet}T_{\bullet}(\mathbb{Q}^{g})\cup\{\infty\} should be read as the bisimplicial pointed set

([p],[q])NqTp(g){},([p],[q])\mapsto N_{q}T_{p}(\mathbb{Q}^{g})\sqcup\{\infty\},

with face and degeneracy operators as defined above.

Remark 2.18.

All automorphism groups in the groupoid Tp(g)T_{p}(\mathbb{Q}^{g}) are trivial, so the canonical map

(22) |NTp(g)|π0(NTp(g))N0(Tp(g))/GLg()|N_{\bullet}T_{p}(\mathbb{Q}^{g})|\to\pi_{0}(N_{\bullet}T_{p}(\mathbb{Q}^{g}))\cong N_{0}(T_{p}(\mathbb{Q}^{g}))/\mathrm{GL}_{g}(\mathbb{Z})

is a weak equivalence for all pp. Therefore the middle space in the zig-zag (21) is also homotopy equivalent to the geometric realization of the simplicial set

[p]N0Tp(g)/GLg(){}.[p]\mapsto N_{0}T_{p}(\mathbb{Q}^{g})/\mathrm{GL}_{g}(\mathbb{Z})\cup\{\infty\}.

We need both simplicial directions for the map to Grg(BK())\operatorname{Gr}_{g}(BK(\mathbb{Z})) though.

We first explain how to construct the map |NT(g){}|Grg(BK())|N_{\bullet}T_{\bullet}(\mathbb{Q}^{g})\cup\{\infty\}|\to\operatorname{Gr}_{g}(BK(\mathbb{Z})) in the zig-zag (21). Here it is natural to look for a map of bisimplicial sets, with (p,q)(p,q)-simplices

NqTp(g){}Grg(NqSp(Proj)).N_{q}T_{p}(\mathbb{Q}^{g})\cup\{\infty\}\to\operatorname{Gr}_{g}(N_{q}S_{p}(\mathrm{Proj}_{\mathbb{Z}})).

We explain how this works for q=0q=0, starting with (=ApA0,<)N0Tp(V)(\emptyset=A_{p}\subset\dots\subset A_{0},<)\in N_{0}T_{p}(V).

The subsets AiV{0}A_{i}\subset V^{\vee}\setminus\{0\} give rise to canonical maps

V\displaystyle V Ai\displaystyle\to\mathbb{Q}^{A_{i}}
v\displaystyle v (ψψ(v)),\displaystyle\mapsto(\psi\mapsto\psi(v)),

which by assumption is injective for i=0i=0. For V=gV=\mathbb{Q}^{g} we shall make use of the restrictions

ggAi\mathbb{Z}^{g}\hookrightarrow\mathbb{Q}^{g}\to\mathbb{Q}^{A_{i}}

in the following. For 0ijp0\leq i\leq j\leq p, define a free \mathbb{Z}-module Pi,jP_{i,j} as the image of the composition

Ker(gAj)gAi.\mathrm{Ker}(\mathbb{Z}^{g}\to\mathbb{Q}^{A_{j}})\hookrightarrow\mathbb{Z}^{g}\to\mathbb{Q}^{A_{i}}.

These modules fit into a triangular diagram of the form (14), forming an object of Sp(Proj)S_{p}(\mathrm{Proj}_{\mathbb{Z}}). Let us also remark that for j=pj=p, the submodule Pi,pAiP_{i,p}\subset\mathbb{Q}^{A_{i}} agrees with the image of the second map gAi\mathbb{Z}^{g}\to\mathbb{Q}^{A_{i}}, which for i=0i=0 is injective. Hence we obtain a canonical isomorphism gP0,p\mathbb{Z}^{g}\to P_{0,p}. We have constructed a map of sets

(23) N0Tp(g)Fg(N0Sp(Proj))Fg1(N0Sp(Proj)),N_{0}T_{p}(\mathbb{Q}^{g})\to F_{g}(N_{0}S_{p}(\mathrm{Proj}_{\mathbb{Z}}))\setminus F_{g-1}(N_{0}S_{p}(\mathrm{Proj}_{\mathbb{Z}})),

which is easily extended to a map of bisimplicial pointed sets

NT(g){}Grg(NS(Proj)),N_{\bullet}T_{\bullet}(\mathbb{Q}^{g})\cup\{\infty\}\to\operatorname{Gr}_{g}(N_{\bullet}S_{\bullet}(\mathrm{Proj}_{\mathbb{Z}})),

which in turn realizes to the desired map of spaces

|NT(g){}|Grg(BK()).|N_{\bullet}T_{\bullet}(\mathbb{Q}^{g})\cup\{\infty\}|\to\operatorname{Gr}_{g}(BK(\mathbb{Z})).
Proposition 2.19.

This map |NT(g){}|Grg(BK())|N_{\bullet}T_{\bullet}(\mathbb{Q}^{g})\cup\{\infty\}|\to\operatorname{Gr}_{g}(BK(\mathbb{Z})) induces an isomorphism in integral homology.

Proof.

The simplicial set N0T(V){}N_{0}T_{\bullet}(V)\cup\{\infty\} has a cubical structure (similar in nature to a space MGrM_{\mathrm{Gr}} which we will define later), as follows. For each pair (A,<)(A,<) consisting of a finite subset AV{0}A\subset V^{\vee}\setminus\{0\} and a total order << on AA, the set of elements (ApA0,<)N0Tp(V){}(A_{p}\subset\dots\subset A_{0},<^{\prime})\in N_{0}T_{p}(V)\cup\{\infty\} for which A0AA_{0}\subset A and <<^{\prime} is the restriction of <<, assemble as pp varies to a map of simplicial sets

(Δ1)A(A,<)N0T(V){}.(\Delta^{1}_{\bullet})^{A}\xrightarrow{(A,<)}N_{0}T_{\bullet}(V)\cup\{\infty\}.

This map lands in the basepoint \infty unless AA spans VV^{\vee}, in which case the induced map of geometric realizations

(24) |Δ1|A|(A,<)||N0T(V)|{}|\Delta^{1}_{\bullet}|^{A}\xrightarrow{|(A,<)|}|N_{0}T_{\bullet}(V)|\cup\{\infty\}

is injective when restricted to (Δ1Δ1)A(\Delta^{1}\setminus\partial\Delta^{1})^{A}. As (A,<)(A,<) varies over all possible (totally ordered) finite spanning sets of VV^{\vee}, the maps (24) together with the basepoint \infty form a CW structure on the space |N0T(V){}||N_{0}T_{\bullet}(V)\cup\{\infty\}|. It follows that the reduced homology of |N0T(V){}||N_{0}T_{\bullet}(V)\cup\{\infty\}| may be computed by a chain complex defined combinatorially by generators [ϕ0,,ϕp](V{0})p+1[\phi_{0},\dots,\phi_{p}]\in(V^{\vee}\setminus\{0\})^{p+1} for varying pp, subject to the relation that [ϕ0,,ϕp]=0[\phi_{0},\dots,\phi_{p}]=0 unless the ϕi\phi_{i} span VV^{\vee}. (Notice that we do not impose any relations between generators that differ by the action of the symmetric group Sp+1S_{p+1} on the set of ordered (p+1)(p+1)-tuples.) The boundary map is given by

[ϕ0,,ϕp]=i=0p(1)i[ϕ0,,ϕi^,,ϕp].\partial[\phi_{0},\dots,\phi_{p}]=\sum_{i=0}^{p}(-1)^{i}[\phi_{0},\dots,\widehat{\phi_{i}},\dots,\phi_{p}].

For V=gV=\mathbb{Q}^{g}, this chain complex is in fact a well known free resolution of Stg()\operatorname{St}_{g}(\mathbb{Q}) as a module over GLg()\mathrm{GL}_{g}(\mathbb{Q}) due to Lee and Szczarba [LS76, Section 3], and we deduce

H~s(|N0T(g){}|){Stg()for p=g0otherwise.\widetilde{H}_{s}(|N_{0}T_{\bullet}(\mathbb{Q}^{g})\cup\{\infty\}|)\cong\begin{cases}\operatorname{St}_{g}(\mathbb{Q})&\text{for $p=g$}\\ 0&\text{otherwise.}\end{cases}

The Steinberg module Stg()\operatorname{St}_{g}(\mathbb{Q}) is also the (reduced) homology of NS~g(Proj){}N_{\bullet}\widetilde{S}^{g}_{\bullet}(\mathrm{Proj}_{\mathbb{Z}})\cup\{\infty\}, the bisimplicial set from the proof of Proposition 2.15 above. It is not hard to compare these two instances of the Steinberg module. Indeed, the map (23) was constructed by a recipe in which the “upper right” module P0,pP_{0,p} in the triangular diagram came with a preferred isomorphism gP0,p\mathbb{Z}^{g}\to P_{0,p}, and this provides a lift to

N0T(g){}N0S~g(Proj){}.N_{0}T_{\bullet}(\mathbb{Q}^{g})\cup\{\infty\}\to N_{0}\widetilde{S}^{g}_{\bullet}(\mathrm{Proj}_{\mathbb{Z}})\cup\{\infty\}.

On reduced homology, the composition

|N0T(g){}||N0S~g(Proj){}||NS~g(Proj){}||N_{0}T_{\bullet}(\mathbb{Q}^{g})\cup\{\infty\}|\to|N_{0}\widetilde{S}^{g}_{\bullet}(\mathrm{Proj}_{\mathbb{Z}})\cup\{\infty\}|\to|N_{\bullet}\widetilde{S}^{g}_{\bullet}(\mathrm{Proj}_{\mathbb{Z}})\cup\{\infty\}|

therefore induces a map between two modules isomorphic to Stg()\operatorname{St}_{g}(\mathbb{Q}). It is the same map as considered by Lee and Szczarba and hence an isomorphism.

By a spectral sequence argument, it follows that the map

|NT(g){}||NSg(Proj)|{}|N_{\bullet}T_{\bullet}(\mathbb{Q}^{g})\cup\{\infty\}|\to|N_{\bullet}S^{g}_{\bullet}(\mathrm{Proj}_{\mathbb{Z}})|\cup\{\infty\}

also induces an isomorphism on homology. ∎

Next we define the map to Grg(Atrop{})=(Pg/GLg()){}\operatorname{Gr}_{g}(A_{\infty}^{\mathrm{trop}}\cup\{\infty\})=(P_{g}/\mathrm{GL}_{g}(\mathbb{Z}))\cup\{\infty\} in (21) as a composition

|NT(g){}||N0T(g)/GLg(){}|Grg(Atrop{}),|N_{\bullet}T_{\bullet}(\mathbb{Q}^{g})\cup\{\infty\}|\xrightarrow{\simeq}|N_{0}T_{\bullet}(\mathbb{Q}^{g})/\mathrm{GL}_{g}(\mathbb{Z})\cup\{\infty\}|\to\operatorname{Gr}_{g}(A_{\infty}^{\mathrm{trop}}\cup\{\infty\}),

where the first map is induced by (22) above, and we must define the second map. To an object (=ApA0,<)N0Tp(g)(\emptyset=A_{p}\subset\dots\subset A_{0},<)\in N_{0}T_{p}(\mathbb{Q}^{g}) we associate a map Δg(Pg/GLg()){}\Delta^{g}\to(P_{g}/\mathrm{GL}_{g}(\mathbb{Z}))\cup\{\infty\}, constructed in the following way. For (=ApA0,<)N0Tp(n)(\emptyset=A_{p}\subset\dots\subset A_{0},<)\in N_{0}T_{p}(\mathbb{Q}^{n}) consider the map

(25) (ΔpΔp)\displaystyle(\Delta^{p}\setminus\partial\Delta^{p}) Pg\displaystyle\to P_{g}
(0<s1<<sp<1)\displaystyle(0<s_{1}<\dots<s_{p}<1) i=1p(si)ψAi1Aiψ2,\displaystyle\mapsto\sum_{i=1}^{p}\ell(s_{i})\sum_{\psi\in A_{i-1}\setminus A_{i}}\psi^{2},

where ψ2Pgrt\psi^{2}\in P_{g}^{\mathrm{rt}} denotes the rank-1 quadratic form v(ψ(v))2v\mapsto(\psi(v))^{2}, and

(26) :(0,1)(0,)\ell\colon(0,1)\to(0,\infty)

is any diffeomorphism, for example slog(s)s\mapsto-\log(s) or ss/(1s)s\mapsto s/(1-s).

Letting (ApA0,<)N0Tp(g)(A_{p}\subset\dots\subset A_{0},<)\in N_{0}T_{p}(\mathbb{Q}^{g}) vary, these maps assemble to a GLg()\mathrm{GL}_{g}(\mathbb{Q})-equivariant map

p(ΔpΔp)×N0Tp(g)Pg.\coprod_{p}(\Delta^{p}\setminus\partial\Delta^{p})\times N_{0}T_{p}(\mathbb{Q}^{g})\to P_{g}.

Passing to orbit sets for the action of the subgroup GLg()\mathrm{GL}_{g}(\mathbb{Z}), we obtain a map which extends as in the following diagram

(27) p(ΔpΔp)×N0Tp(g)/GLg(){\coprod_{p}(\Delta^{p}\setminus\partial\Delta^{p})\times N_{0}T_{p}(\mathbb{Q}^{g})/\mathrm{GL}_{g}(\mathbb{Z})}Pg/GLg(){P_{g}/\mathrm{GL}_{g}(\mathbb{Z})}|N0T(g)/GLg(){}|{{|N_{0}T_{\bullet}(\mathbb{Q}^{g})/\mathrm{GL}_{g}(\mathbb{Z})\cup\{\infty\}|}}(Pg/GLg()){}.{(P_{g}/\mathrm{GL}_{g}(\mathbb{Z}))\cup\{\infty\}.}
Proposition 2.20.

The bottom horizontal map in (27) induces an isomorphism in reduced rational homology.

We remark that a map similar to the bottom horizontal map in (27) appeared in [Ash94, p. 333] where it is credited to L. Rudolph.

Proof.

The homology of the domain is isomorphic to Eg,1Q=H(GLg();Stg){}^{Q}\!E^{1}_{g,*}=H_{*}(\mathrm{GL}_{g}(\mathbb{Z});\operatorname{St}_{g}\otimes\mathbb{Q}), as is the homology of the codomain by duality (see §2.1.1). Therefore the rational homology groups of domain and codomain are abstractly isomorphic. Since the homology groups are also finite-dimensional, it suffices to construct a one-sided inverse to the map.

To produce a one-sided inverse (on the level of rational homology), the strategy will be to choose a suitable “simplicial structure” on Pg/GLg(){}P_{g}/\mathrm{GL}_{g}(\mathbb{Z})\cup\{\infty\}, namely a simplicial pointed set XX_{\bullet} with finitely many non-degenerate simplices, and a homotopy equivalence |X|Pg/GLg(){}|X_{\bullet}|\to P_{g}/\mathrm{GL}_{g}(\mathbb{Z})\cup\{\infty\}, and then defining the one-sided inverse on the simplicial chains of XX_{\bullet} by sending each non-degenerate non-basepoint simplex to some preferred lift in N0T(g)/GLg(){}N_{0}T_{\bullet}(\mathbb{Q}^{g})/\mathrm{GL}_{g}(\mathbb{Z})\cup\{\infty\}. We will not quite succeed in doing exactly this, but can achieve that simplices of XX_{\bullet} come with a finite set of specified lifts. The one-sided inverse is given by averaging those lifts.

To construct the simplicial set XX_{\bullet}, we first choose an admissible decomposition of PgrtP_{g}^{\mathrm{rt}} whose rays are all rank 11, for instance the perfect cone decomposition. Each cone σ\sigma is then the convex hull of its extremal rays 0Pgrt\mathbb{R}_{\geq 0}\to P_{g}^{\mathrm{rt}} which are of the form ttψ2t\mapsto t\psi^{2} for some ψ(g){0}\psi\in(\mathbb{Q}^{g})^{\vee}\setminus\{0\}. Up to scaling by a non-zero rational number we can arrange that ψ\psi is normalized to satisfy ψ(g)=\psi(\mathbb{Z}^{g})=\mathbb{Z}, and for normalized ψ\psi the form ψ2Pgrt\psi^{2}\in P_{g}^{\mathrm{rt}} is uniquely determined by the projective class [ψ]((g))[\psi]\in\mathbb{P}((\mathbb{Q}^{g})^{\vee}), because ψ\psi itself is determined up to a sign by the normalization condition.

Any cone in the cone decomposition is then the image of a map of the form

0p\displaystyle\mathbb{R}_{\geq 0}^{p} Pgrt\displaystyle\to P_{g}^{\mathrm{rt}}
(t1,,tp)\displaystyle(t_{1},\dots,t_{p}) i=1ptiψi2\displaystyle\mapsto\sum_{i=1}^{p}t_{i}\psi_{i}^{2}

for some normalized ψi(g){0}\psi_{i}\in(\mathbb{Q}^{g})^{\vee}\setminus\{0\}. The map need not be injective unless the cone is simplicial, but it will always be a proper homotopy equivalence onto the cone (the inverse image of any point in the cone is a compact and convex subset of the octant). If the ψi\psi_{i} span (g)(\mathbb{Q}^{g})^{\vee}, then the open cone has image in PgP_{g} and is also the image of the map

(Δ1Δ1)p\displaystyle(\Delta^{1}\setminus\partial\Delta^{1})^{p} Pg\displaystyle\to P_{g}
(s1,,sp)\displaystyle(s_{1},\dots,s_{p}) i=1p(si)ψi2,\displaystyle\mapsto\sum_{i=1}^{p}\ell(s_{i})\psi_{i}^{2},

where \ell is as in (26), analogous to (25) but without inequalities among the sis_{i}. The admissible decomposition gives finitely many GLg()\mathrm{GL}_{g}(\mathbb{Z})-orbits of such maps, whose image in Pg/GLg()P_{g}/\mathrm{GL}_{g}(\mathbb{Z}) forms a decomposition into open cones (which may or may not be simplicial). After 1-point compactifying, we obtain finitely many maps of the form

(28) |Δ1|p(Pg/GLg()){},|\Delta^{1}_{\bullet}|^{p}\to(P_{g}/\mathrm{GL}_{g}(\mathbb{Z}))\cup\{\infty\},

one for each GLg()\mathrm{GL}_{g}(\mathbb{Z})-orbit of cones spanned by pp many rays ttψi2t\mapsto t\psi_{i}^{2}, with the property that the ψi\psi_{i} span (g)(\mathbb{Q}^{g})^{\vee}. The image of each such map is the closure of the corresponding cone, but the map need not be injective. However, the inverse image of a point in the cone is contractible. If we pre-compose with the geometric realization of any of the top-dimensional simplices Δp(Δ1)p\Delta^{p}_{\bullet}\to(\Delta^{1}_{\bullet})^{p} then we obtain p!p! many maps of the form (25) for each cone.

The maps ΔpPg/GLg(){}\Delta^{p}\to P_{g}/\mathrm{GL}_{g}(\mathbb{Z})\cup\{\infty\} arising this way, p!p! many for each cone in the decomposition with pp many extremal rays, can be regarded as elements of the pointed set Sinp((Pg/GLg(){})\mathrm{Sin}_{p}((P_{g}/\mathrm{GL}_{g}(\mathbb{Z})\cup\{\infty\}). Together with all iterated faces and degeneracies of these elements, and with the basepoint, they form a simplicial subset

XSin((Pg/GLg(){})X_{\bullet}\subset\mathrm{Sin}_{\bullet}((P_{g}/\mathrm{GL}_{g}(\mathbb{Z})\cup\{\infty\})

with the property that the induced map

|X|Pg/GLg(){}|X_{\bullet}|\to P_{g}/\mathrm{GL}_{g}(\mathbb{Z})\cup\{\infty\}

has contractible fibers and that XX has only finitely many non-degenerate simplices. It follows from the Vietoris–Begle theorem that this map induces an isomorphism on homology. The simplicial set XX_{\bullet} has only finitely many non-degenerate simplices, on each of which the map is of the form (25) and therefore agrees with the map |N0T(g){}|Pg/GLg(){}|N_{0}T_{\bullet}(\mathbb{Q}^{g})\cup\{\infty\}|\to P_{g}/\mathrm{GL}_{g}(\mathbb{Z})\cup\{\infty\} restricted to some pp-simplex.

If we write C~(X)=C(X,{};)\widetilde{C}_{*}(X)=C_{*}(X_{\bullet},\{\infty\};\mathbb{Q}) for the normalized rational chains relative to the base point, then we obtain a quasi-isomorphism

C~(X)C~(Sin(Pg/GLg(){}))C~sing(Pg/GLg(){};)\widetilde{C}_{*}(X_{\bullet})\stackrel{{\scriptstyle\simeq}}{{\hookrightarrow}}\widetilde{C}_{*}(\mathrm{Sin}_{\bullet}(P_{g}/\mathrm{GL}_{g}(\mathbb{Z})\cup\{\infty\}))\simeq\widetilde{C}_{*}^{\mathrm{sing}}(P_{g}/\mathrm{GL}_{g}(\mathbb{Z})\cup\{\infty\};\mathbb{Q})

from a finite chain complex C~(X)\widetilde{C}_{*}(X_{\bullet}) similar in spirit to the “Voronoi complexes” of [EVGS13]. The main difference is that each cone in the decomposition, of geometric dimension dd and spanned by pp many extremal rays, gives rise to one generator of homological degree dd in the Voronoi complex while in our complex it gives rise to p!p! many generators of homological degree pp and also some lower-dimensional generators.

Unfortunately, this does not quite identify XX_{\bullet} with a simplicial subset of N0T(g){}N_{0}T_{\bullet}(\mathbb{Q}^{g})\cup\{\infty\}, for two reasons. Firstly, the extremal rays of the cones in a cone decomposition are of the form ttψ2t\mapsto t\psi^{2} for some covector ψ(g)\psi\in(\mathbb{Q}^{g})^{\vee} but ψ\psi is not quite canonically determined by the ray, only up to a sign. Secondly, elements of N0Tp(g)N_{0}T_{p}(\mathbb{Q}^{g}) involve a total order << which we cannot canonically produce from a pp-simplex of XX_{\bullet}. Altogether, we have explained a recipe by which a non-degenerate non-basepoint σXp\sigma\in X_{p} lifts in 2pp!2^{p}p! many ways to an element σ~N0Tp(g)/GLg()\widetilde{\sigma}\in N_{0}T_{p}(\mathbb{Q}^{g})/\mathrm{GL}_{g}(\mathbb{Z}). This is sufficient for defining a one-sided inverse on the level of rational homology though: let the diagonal map in the diagram

C~sing(|N0T(g)/GLg(){}|;){\widetilde{C}_{*}^{\mathrm{sing}}(|N_{0}T_{\bullet}(\mathbb{Q}^{g})/\mathrm{GL}_{g}(\mathbb{Z})\cup\{\infty\}|;\mathbb{Q})}C~(X){\widetilde{C}_{*}(X_{\bullet})}C~sing(Pg/GLg(){};){\widetilde{C}_{*}^{\mathrm{sing}}(P_{g}/\mathrm{GL}_{g}(\mathbb{Z})\cup\{\infty\};\mathbb{Q})}\scriptstyle{\simeq}

be defined by sending a non-degenerate non-basepoint simplex of XX_{\bullet} to the average of the 2pp!2^{p}p! many lifts to N0Tp(g)N_{0}T_{p}(\mathbb{Q}^{g}) that we have explained. ∎

Remark 2.21.

The choice of diffeomorphism \ell in (26) may be chosen either to be an increasing or decreasing function. This gives two different isomorphisms between Es,t1T{}^{T}\!E^{1}_{s,t} and Es,t1Q{}^{Q}\!E^{1}_{s,t}, differing by the automorphism of EQ{}^{Q}\!E^{*} induced by dualizing projective modules (and reflecting the triangular diagrams (14)). The two comparison isomorphisms differ by a sign on E1,01QE1,01T{}^{Q}\!E^{1}_{1,0}\cong\mathbb{Q}\cong{}^{T}\!E^{1}_{1,0}.

Remark 2.22.

Orthogonal direct sum of symmetric forms induces a map Pg×PhPg+hP_{g}\times P_{h}\to P_{g+h}. Passing to orbits leads to a map

(Pg/GLg())×(Ph/GLh())(Pg+h/GLg+h())(P_{g}/\mathrm{GL}_{g}(\mathbb{Z}))\times(P_{h}/\mathrm{GL}_{h}(\mathbb{Z}))\to(P_{g+h}/\mathrm{GL}_{g+h}(\mathbb{Z}))

which is proper. Therefore it induces a map of one-point compactifications, giving a product on E,1T{}^{T}\!E^{1}_{*,*}, the E1E^{1}-page of the tropical spectral sequence. The isomorphism

E1TE1Q{}^{T}\!E^{1}\xrightarrow{\cong}{}^{Q}\!E^{1}

constructed above is an isomorphism of bigraded algebras with respect to this product on E,1T{}^{T}\!E^{1}_{*,*} and the product on E,1Q{}^{Q}\!E^{1}_{*,*} constructed in Section 3. There, we also produce a coproduct, of which it seems more difficult to give a simple interpretation in ET{}^{T}\!E.

2.5. Canonical forms for GLn\mathrm{GL}_{n}

The space PgP_{g} of positive-definite real symmetric matrices of rank gg is equipped with a right action of GLg()\mathrm{GL}_{g}(\mathbb{R}) given by XgTXgX\mapsto g^{T}Xg. Classical invariant theory provides a differential nn-form for every n1n\geq 1

ωXn=tr((X1dX)n)\omega^{n}_{X}=\operatorname{tr}((X^{-1}dX)^{n})

which is invariant under the action of GLg()\mathrm{GL}_{g}(\mathbb{R}). One shows that ωn\omega^{n} vanishes unless n1(mod4)n\equiv 1\pmod{4}; the case n=1n=1 plays very little role in what follows. The ωn\omega^{n} have a number of useful properties, including compatibility with block direct sums of matrices:

(29) ωX1X2n=ωX1n+ωX2n\omega^{n}_{X_{1}\oplus X_{2}}=\omega^{n}_{X_{1}}+\omega^{n}_{X_{2}}

and, in the case when h>1h>1 is odd, the form ω2h1\omega^{2h-1} has the vanishing property

(30) ωX2h1=0 if X has rank g<h.\omega^{2h-1}_{X}=0\qquad\hbox{ if }X\hbox{ has rank }g<h\ .

By invariance, the forms ωX4k+1\omega^{4k+1}_{X} for k>1k>1 define differential forms on the locally symmetric space Pg/GLg()P_{g}/\mathrm{GL}_{g}(\mathbb{Z}). Borel showed that the graded exterior algebra they generate is isomorphic to its stable cohomology:

k1ω4k+1limgH(Pg/GLg();).\bigwedge\bigoplus_{k\geq 1}\omega^{4k+1}\mathbb{R}\ \overset{\sim}{\longrightarrow}\ \varprojlim_{g}H^{*}(P_{g}/\mathrm{GL}_{g}(\mathbb{Z});\mathbb{R})\ .

Both sides of this equation have a Hopf algebra structure such that the (indecomposable) generators ω4k+1\omega^{4k+1} are primitive. This follows from (29). By taking primitives, Borel deduced that Kn()K_{n}(\mathbb{Z})\otimes\mathbb{Q}\cong\mathbb{Q} for all n=4k+1n=4k+1 with k>1k>1, and vanishes for all other n>0n>0.

We will need much more precise results about the cohomology of GLn()\mathrm{GL}_{n}(\mathbb{Z}) in the unstable range. For this, let g>1g>1 be odd and let Ω(g)\Omega^{*}(g) denote the graded exterior algebra generated by ω5,,ω2g1\omega^{5},\ldots,\omega^{2g-1}, which are the non-vanishing forms on Pg/GLg().P_{g}/\mathrm{GL}_{g}(\mathbb{Z}). It is a direct sum

Ω(g)=Ωc(g)Ωnc(g)\Omega^{*}(g)=\Omega^{*}_{c}(g)\oplus\Omega^{*}_{nc}(g)

where Ωc(g)\Omega^{*}_{c}(g) is the graded vector space of forms ‘of compact type’ given by the ideal generated by ω2g1\omega^{2g-1}, and Ωnc(g)\Omega^{*}_{nc}(g) is the graded algebra generated by ω5,,ω2g5\omega^{5},\ldots,\omega^{2g-5}. The Hodge star operator interchanges these two spaces. In [Bro23] it was shown that every element in Ωc(g)\Omega_{c}(g) defines a unique compactly supported cohomology class, giving two injective maps

(31) Ωnc(g)\displaystyle\Omega^{*}_{nc}(g) \displaystyle\longrightarrow H(Ph/GLh();) for all hg,\displaystyle H^{*}(P_{h}/\mathrm{GL}_{h}(\mathbb{Z});\mathbb{R})\qquad\hbox{ for all }h\geq g,
(32) Ωc(g)[1]\displaystyle\Omega^{*}_{c}(g)[-1] \displaystyle\longrightarrow Hc(Pg/GLg();),\displaystyle H_{c}^{*}(P_{g}/\mathrm{GL}_{g}(\mathbb{Z});\mathbb{R})\ ,

the first of which is a map of graded algebras, the second only a map of graded vector spaces. By taking the limit of the first injection, one obtains a stronger version of Borel’s theorem (which states that (31) is injective in degrees g/4\leq g/4 when h=gh=g.) However, it is the second injective map (32) which is involved in a key definition.

Definition 2.23.

Set Ωc(g)=0\Omega_{c}(g)=0 for gg even. Denote by

Ωc=g>1Ωc(g)\Omega^{*}_{c}=\bigoplus_{g>1}\Omega_{c}^{*}(g)

the \mathbb{Q} vector space spanned by the canonical forms of compact type, bigraded as follows. The degree and genus are given on generators by

deg(ω4i1+1ω4ik+1)\displaystyle\operatorname{deg}(\omega^{4i_{1}+1}\wedge\cdots\wedge\omega^{4i_{k}+1}) =\displaystyle= Σj=1k(4ij+1)\displaystyle\textstyle\Sigma_{j=1}^{k}(4i_{j}+1)
g(ω4i1+1ω4ik+1)\displaystyle\operatorname{g}(\omega^{4i_{1}+1}\wedge\cdots\wedge\omega^{4i_{k}+1}) =\displaystyle= 2ik+1\displaystyle 2i_{k}+1

for integers 0<i1<<ik.0<i_{1}<\cdots<i_{k}. Then a generator ω\omega is in bidegree (g(ω),deg(ω)g).(\operatorname{g}(\omega),\operatorname{deg}(\omega)\!-\!\operatorname{g}).

Although the bigraded vector space Ωc\Omega^{*}_{c} is formally isomorphic to the space of elements of positive degree in the graded exterior algebra generated by the {ω4k+1k1}\{\omega^{4k+1}\mid k\geq 1\}, it is advisable not to confuse the two. The former is a bigraded vector space, while the latter is the positive degree elements of a Hopf algebra. Indeed, we shall prove that the image of Ωc\Omega^{*}_{c} is primitive with respect to the coproduct we shall define, which is not the case for non-trivial products of ω4k+1\omega^{4k+1} in the stable cohomology of the general linear group. The precise relationship between these two coproducts is via the Quillen spectral sequence, explained below.

2.5.1. Canonical tensor algebra

Denote the tensor algebra on a bigraded vector space VV by

T(V)=n0Vn.T(V)=\bigoplus_{n\geq 0}V^{\otimes n}\ .

It is a connected, bigraded Hopf algebra with non-commutative product given by the tensor product, and (graded) cocommutative coproduct Δ:T(V)T(V)T(V)\Delta\colon T(V)\rightarrow T(V)\otimes T(V) dual to the shuffle product, with respect to which the elements of VV are primitive. We shall denote elements v1vnVnT(V)v_{1}\otimes\ldots\otimes v_{n}\in V^{\otimes n}\subset T(V) using the bar notation [v1||vn][v_{1}|\ldots|v_{n}].

Let Ωc[1]\Omega^{\ast}_{c}[-1] denote the bigraded \mathbb{Q}-vector space in which degree is shifted by 11. So ω4i+1\omega^{4i+1} has genus 2i+12i+1 and degree (4i+1)+1(4i+1)+1. The tensor algebra T(Ωc[1])T(\Omega^{*}_{c}[-1]) is a non-commutative, (graded) cocommutative Hopf algebra. It is again bigraded by genus and degree minus genus, where [v1||vn][v_{1}|\ldots|v_{n}] has degree deg(vi)\sum\mathrm{deg}(v_{i}) and genus g(vi)\sum\mathrm{g}(v_{i}). Note that it has an additional grading by length of tensors, but only the associated filtration will play a role.

3. Coproducts and filtrations in the Waldhausen construction

Recall that the rational homology of an infinite loop space such as BK()BK(\mathbb{Z}) is naturally a graded commutative and cocommutative Hopf algebra, with coproduct given by push-forward under the diagonal. However, this coproduct does not respect the filtration on BK()BK(\mathbb{Z}) and hence does not give rise to a coproduct on the pages of the Quillen spectral sequence.

Our goal in this section is to construct a filtered coproduct on BK()BK(\mathbb{Z}). We do so via the simplicial operation of edgewise subdivision from [BHM93, Lemma 1.1]. The resulting coproduct induces a bigraded associative (but not cocommutative) coproduct on each page of the Quillen spectral sequence. It is homotopic to the diagonal and hence induces the usual commutative coproduct on the abutment.

3.1. A filtered coproduct from edgewise subdivision

The goal now is to define a map

BK()BK()×BK()BK(\mathbb{Z})\to BK(\mathbb{Z})\times BK(\mathbb{Z})

that respects the filtration (16), i.e., with the property that the restriction to the nnth filtration Fn=FnBK()F_{n}=F_{n}BK(\mathbb{Z}) factors as

Fna+b=nFa×FbF_{n}\to\bigcup_{a+b=n}F_{a}\times F_{b}

so that it induces a coproduct on the associated relative homology spectral sequence. We will use the edgewise subdivision of [BHM93, Section 1], in turn inspired by [Seg73]. We recall the definition.

Definition 3.1.

For any category 𝒮\mathcal{S}, given a simplicial object X:Δop𝒮X\colon\Delta^{\mathrm{op}}\to\mathcal{S}, the edgewise subdivision es(X)\mathrm{es}(X) of XX is the simplicial object in 𝒮\mathcal{S}

ΔΔ𝑋𝒮\Delta\to\Delta\xrightarrow{X}\mathcal{S}

where the functor ΔΔ\Delta\to\Delta is

[p][p][p].[p]\mapsto[p]\sqcup[p].

Here, the set [p][p][p]\sqcup[p] is linearly ordered by concatenation.

In other words, we identify [p][p][1]×[p][p]\sqcup[p]\cong[1]\times[p] with the lexicographic ordering.

Remark 3.2.

Note, in particular, that esp(X)=X2p+1\mathrm{es}_{p}(X)=X_{2p+1} for each p0.p\geq 0.

For a simplicial set X:Δop𝖲𝖾𝗍X\colon\Delta^{\mathrm{op}}\to\mathsf{Set}, there is a natural homeomorphism

|es(X)|ϕX|X|.|\mathrm{es}(X)|\xrightarrow{\phi_{X}}|X|.

It is the unique natural transformation that is affine on simplices and sends each vertex xes0(X)=X1x\in\mathrm{es}_{0}(X)=X_{1} to the mid-point of the corresponding edge or vertex (degenerate 1-simplex) in |X||X|. To check that this is a homeomorphism it suffices to consider the case where X=Δ(,[n])X=\Delta(-,[n]) is a simplex, since both sides preserve colimits. See [BHM93, Lemma 1.1].

We now consider how the (inverse of the) homeomorphism ϕX\phi_{X} interacts with filtrations.

Lemma 3.3.

Let XX_{\bullet} be a filtered simplicial set, i.e., a simplicial set XX and a sequence of subsets F0XpF1XpF_{0}X_{p}\subset F_{1}X_{p}\subset\dots for each pp such that each FnXF_{n}X_{\bullet} forms a simplicial subset, i.e. diFnXp+1FnXpd_{i}F_{n}X_{p+1}\subset F_{n}X_{p} and siFnXpFnXp+1s_{i}F_{n}X_{p}\subset F_{n}X_{p+1} for all pp and all ii. For any filtered simplicial space, give the geometric realization |X||X| the induced filtration: Fn|X|F_{n}|X| is the image of the realization of the inclusion |FnX||X||F_{n}X|\to|X|. Finally, give the simplicial set es(X)\mathrm{es}_{\bullet}(X) the filtration with Fnesp(X)=esp(FnX)=FnX2p+1F_{n}\mathrm{es}_{p}(X)=\mathrm{es}_{p}(F_{n}X_{\bullet})=F_{n}X_{2p+1}.

Then the inverse homeomorphism

|X|ϕX1|es(X)||X|\xrightarrow{\phi_{X}^{-1}}|\mathrm{es}(X)|

is a filtered map: it satisfies ϕX1(Fn|X|)Fn|es(X)|\phi_{X}^{-1}(F_{n}|X|)\subset F_{n}|\mathrm{es}(X)|.

Proof.

Let f:Xpf\colon X_{p}\to\mathbb{N} be the function such that f(σ)=nf(\sigma)=n if σFnXpFn1Xp\sigma\in F_{n}X_{p}\setminus F_{n-1}X_{p}. Then each FnXF_{n}X_{\bullet} being a simplicial subset implies that f(θσ)f(σ)f(\theta^{*}\sigma)\leq f(\sigma) for any morphism θ:[p][q]\theta\colon[p]\to[q] in Δ\Delta and any σXq\sigma\in X_{q}, but then the simplicial identity disi(σ)=σd_{i}\circ s_{i}(\sigma)=\sigma implies f(siσ)=f(σ)f(s_{i}\sigma)=f(\sigma). In other words, the function ff is determined by its values on non-degenerate simplices.

Now a point x|X|x\in|X| will be in the image of the canonical injection

{σx}×(ΔpΔp)|X|\{\sigma_{x}\}\times(\Delta^{p}\setminus\partial\Delta^{p})\hookrightarrow|X|

for a unique pp\in\mathbb{N} and a unique non-degenerate σxXp\sigma_{x}\in X_{p}. Then ϕX1:|X||es(X)|\phi_{X}^{-1}\colon|X|\xrightarrow{\sim}|\mathrm{es}(X)| sends xx to a point in the image of the canonical injection

{τx}×(ΔqΔq)|es(X)|\{\tau_{x}\}\times(\Delta^{q}\setminus\partial\Delta^{q})\hookrightarrow|\mathrm{es}(X)|

for a unique q{0,,p}q\in\{0,\dots,p\} and a unique τxesq(X)=X2q+1\tau_{x}\in\mathrm{es}_{q}(X)=X_{2q+1}.

By naturality of the homeomorphism, there must exist a morphism θx:[p][2q+1]\theta_{x}\colon[p]\to[2q+1] in Δ\Delta such that θx(σx)=τx\theta_{x}^{*}(\sigma_{x})=\tau_{x}. This implies f(τx)f(σx)f(\tau_{x})\leq f(\sigma_{x}), and hence

f(ϕX1(x))f(x)f(\phi_{X}^{-1}(x))\leq f(x)

as desired. ∎

Let us briefly discuss edgewise subdivision of bisimplicial sets X:Δop×Δop𝖲𝖾𝗍X\colon\Delta^{\mathrm{op}}\times\Delta^{\mathrm{op}}\to\mathsf{Set}, for which we write Xp,q=X([p],[q])X_{p,q}=X([p],[q]). The notion itself is symmetric in pp and qq, but for the applications we have in mind the two simplicial directions will play quite different roles. In particular, when 𝒞\mathcal{C} is a Waldhausen category we will consider the bisimplicial set with

Xp,q=NqSp(𝒞)X_{p,q}=N_{q}S_{p}(\mathcal{C})

and use the first simplicial direction pp as the “SS_{\bullet}-direction.” In this and other examples we consider, the second simplicial direction will play a more auxiliary role, mainly as a means to encode the simplicial topological space [p]|Xp,|[p]\mapsto|X_{p,\bullet}|, while in the pp-direction we will make more use of the simplicial structure. In particular, we will write

es(X)=es([p]Xp,)\mathrm{es}(X)=\mathrm{es}([p]\mapsto X_{p,\bullet})

for edgewise subdivision in the pp-direction. In other words, on objects [p][p] and [q][q] of Δ\Delta we have

es(X)p,q=X2p+1,q=X([p][p],[q]).\mathrm{es}(X)_{p,q}=X_{2p+1,q}=X([p]\sqcup[p],[q]).
Proposition 3.4.

Let X:Δop×Δop𝖲𝖾𝗍X\colon\Delta^{\mathrm{op}}\times\Delta^{\mathrm{op}}\to\mathsf{Set} be a bisimplicial set. Then |es(X)||\mathrm{es}(X)| is naturally homeomorphic to |X||X|.

Proof.

We have

|X|\displaystyle|X| \displaystyle\cong |[q]([p]Xp,q)||[q]([p]X2p+1,q)||[p]([q]X2p+1,q)||es(X)|,\displaystyle|[q]\mapsto([p]\mapsto X_{p,q})|\cong|[q]\mapsto([p]\mapsto X_{2p+1,q})|\cong|[p]\mapsto([q]\mapsto X_{2p+1,q})|\cong|\mathrm{es}(X)|,

using the fact that edgewise subdivision is a natural homeomorphism for simplicial sets, and that the two different geometric realizations of bisimplicial sets explained above are naturally homeomorphic. All of the other intermediate homeomorphisms are natural as well. ∎

Here is the main definition.

Definition 3.5.

Let 𝒮\mathcal{S} be a category with finite products. Let

X:Δop𝒮X\colon\Delta^{\mathrm{op}}\to\mathcal{S}

be a simplicial object in 𝒮\mathcal{S}, and let es(X)\mathrm{es}(X) be its edgewise subdivision. Define a natural transformation

(33) es(X)X×X\mathrm{es}(X)\Rightarrow X\times X

whose component at [p][p], for each p0p\geq 0,

es(X)p=X2p+1Xp×Xp,\mathrm{es}(X)_{p}=X_{2p+1}\to X_{p}\times X_{p},

is given by the two maps X2p+1XpX_{2p+1}\to X_{p} induced by the two order preserving inclusions

[p][p][p][2p+1][p]\to[p]\sqcup[p]\cong[2p+1]

onto the first p+1p+1, respectively the last p+1p+1, elements.

Remark 3.6.

When 𝒮\mathcal{S} is the category of simplicial sets, X:Δop𝒮X\colon\Delta^{\mathrm{op}}\to\mathcal{S} is a bisimplicial set. Using that geometric realization commutes with finite products, the natural transformation es(X)X×X\mathrm{es}(X)\Rightarrow X\times X defined above leads to a map

(34) |X||es(X)||X|×|X|.|X|\xrightarrow{\approx}|\mathrm{es}(X)|\to|X|\times|X|.

As a map of spaces, this map is not especially interesting by itself—as we will see momentarily, it is naturally homotopic to the diagonal map x(x,x)x\mapsto(x,x) of the space |X||X|. What we will see is that in interesting examples, the bisimplicial set XX comes with a filtration making (34) into a map of filtered spaces and (34) will induce an interesting map of associated gradeds, while the actual diagonal map of |X||X| will not be a filtered map.

Example 3.7.

Our main example comes from the simplicial object X=S(Proj):[p]Sp(Proj)X=S_{\bullet}(\mathrm{Proj}_{\mathbb{Z}})\colon[p]\mapsto S_{p}(\mathrm{Proj}_{\mathbb{Z}}) in small groupoids, for the Waldhausen category Proj\mathrm{Proj}_{\mathbb{Z}} of finitely generated projective \mathbb{Z}-modules (Example 2.11). In this case the functor

esp(S(Proj))Sp(Proj)×Sp(Proj)\mathrm{es}_{p}(S_{\bullet}(\mathrm{Proj}_{\mathbb{Z}}))\to S_{p}(\mathrm{Proj}_{\mathbb{Z}})\times S_{p}(\mathrm{Proj}_{\mathbb{Z}})

is given on objects by

(0V1V2p+1)((0V1Vp),(0Vp+2/Vp+1V2p+1/Vp+1)),(0\to V_{1}\to\cdots\to V_{2p+1})\mapsto((0\to V_{1}\to\cdots\to V_{p}),(0\to V_{p+2}/V_{p+1}\to\cdots\to V_{2p+1}/V_{p+1})),

omitting the chosen quotients Vi/VjV_{i}/V_{j} from the notation.

We recall the filtration on X=S(Proj)X=S_{\bullet}(\mathrm{Proj}_{\mathbb{Z}}) described previously in (16). For integers p,qp,q, an element of Xp,qX_{p,q} is a chain of qq morphisms of triangular diagrams of projective modules of the form (14). The morphisms of triangular diagrams are required to be isomorphisms in each component. Define the rank of such an element to be the rank of the top-right projective \mathbb{Z}-module in any of the q+1q+1 triangular diagrams; this rank is well-defined since the q+1q+1 modules involved are related by isomorphisms. For any nn, the property of having rank at most nn is preserved by all face and degeneracy maps of XX, and thus rank induces a filtered bisimplicial set in which

(FnX)p,q=Xp,qrank1((,n]).(F_{n}X)_{p,q}=X_{p,q}\cap\mathrm{rank}^{-1}((-\infty,n]).

The rank function on simplices of XX induces a rank function on the simplices of es(X)\mathrm{es}(X), as well as a rank function on the simplices of X×XX\times X by additivity: the rank of (x,y)(X×X)p,q=Xp,q×Xp,q(x,y)\in(X\times X)_{p,q}=X_{p,q}\times X_{p,q} is rank(x)+rank(y)\mathrm{rank}(x)+\mathrm{rank}(y). These rank functions are compatible with face and degeneracy maps, inducing filtrations on both es(X)\mathrm{es}(X) and X×XX\times X. We note that (33) respects the filtration, since

(35) rank(V2p+1)rank(Vp)+rank(V2p+1/Vp+1).\mathrm{rank}(V_{2p+1})\geq\mathrm{rank}(V_{p})+\mathrm{rank}(V_{2p+1}/V_{p+1}).

Therefore, the natural transformation es(X)X×X\mathrm{es}(X)\Rightarrow X\times X is a map of filtered bisimplicial sets, and geometric realization yields a map

|X||es(X)||X×X||X|×|X||X|\xrightarrow{\approx}|\mathrm{es}(X)|\longrightarrow|X\times X|\cong|X|\times|X|

of filtered spaces.

In Proposition 3.13, we shall study a product map m:X×XXm\colon X\times X\to X, given by direct sums of triangular diagrams. The observation that the rank of a direct sum of projective modules is the sum of the ranks will imply that mm is compatible with filtrations; see Remark 3.14 below.

Returning briefly to the general case of an arbitrary (unfiltered) simplicial set XX, let us justify the claim above that (34) is homotopic to the diagonal.

Proposition 3.8.

Suppose we are given maps ΦX:|X||X|\Phi_{X}\colon|X|\to|X| for every simplicial set XX that are natural in XX, i.e., they assemble into a natural transformation Φ\Phi. Then Φ\Phi is naturally homotopic to the identity. Precisely, there is a homotopy

|X|×[0,1]|X||X|\times[0,1]\to|X|

from ΦX\Phi_{X} to idX\mathrm{id}_{X} for each XX, and these homotopies are natural in XX.

Proof.

First, for X=Δn=HomΔ(,[n])X=\Delta^{n}=\operatorname{Hom}_{\Delta}(-,[n]) an nn-simplex, one may take a straight line homotopy

|Δn|×[0,1]|Δn||\Delta^{n}|\times[0,1]\to|\Delta^{n}|

from ΦX\Phi_{X} to idX\mathrm{id_{X}}, since Δn\Delta^{n} is a convex subset of n\mathbb{R}^{n}. These straight line homotopies are natural in all morphisms of simplicial sets, in particular face maps and degeneracy maps. An arbitrary simplicial set is a colimit of simplices, so the result follows.∎

Corollary 3.9.

For simplicial sets XX, the geometric realization

(36) |X||es(X)||X|×|X||X|\cong|\mathrm{es}(X)|\to|X|\times|X|

of the morphism in (33) is naturally homotopic to the diagonal map.

Proof.

The natural transformation es(X)X×X\mathrm{es}(X)\Rightarrow X\times X is assembled from two natural transformations es(X)X\mathrm{es}(X)\to X, with the property that the corresponding maps |es(X)||X||\mathrm{es}(X)|\to|X| are homotopic to the identity by Proposition 3.8. Therefore the map |es(X)||X|×|X||\mathrm{es}(X)|\to|X|\times|X| is naturally homotopic to the diagonal.∎

Remark 3.10.

Suppose as in Proposition 3.8, we are given maps ΦX:|X||X|\Phi_{X}\colon|X|\to|X| for every simplicial set XX that are natural in XX. Now if XX is a bisimplicial set, for each p0p\geq 0, let XpX_{p} denote the simplicial set with (Xp)q=Xp,q(X_{p})_{q}=X_{p,q}. Then we obtain a map ΦX:|X||X|\Phi_{X}\colon|X|\to|X| by gluing the maps ΦX:|Xp||Xp|\Phi_{X}\colon|X_{p}|\to|X_{p}| for each pp. Then Proposition 3.8 implies that ΦX\Phi_{X} is also naturally homotopic to the identity.

In particular, for a bisimplicial set XX, the map |X||es(X)||X|×|X||X|\cong|\mathrm{es}(X)|\to|X|\times|X| is again homotopic to the diagonal, naturally in XX.

Proposition 3.11.

Suppose

F0XF1XX:Δop×Δop𝖲𝖾𝗍F_{0}X\subset F_{1}X\subset\cdots\subset X\colon\Delta^{\mathrm{op}}\times\Delta^{\mathrm{op}}\to\mathsf{Set}

is a filtered bisimplicial set. Define a filtration on the bisimplicial set X×XX\times X by

Fs(X×X)p,q=u+vs(FuX×FvX)p,q,F_{s}(X\times X)_{p,q}=\bigcup_{u+v\leq s}(F_{u}X\times F_{v}X)_{p,q},

and similarly for X×X×XX\times X\times X:

Fs(X×X×X)p,q=t+u+vs(FtX×FuX×FvX)p,q.F_{s}(X\times X\times X)_{p,q}=\bigcup_{t+u+v\leq s}(F_{t}X\times F_{u}X\times F_{v}X)_{p,q}.

Let Φ:es(X)X×X\Phi\colon\mathrm{es}(X)\Rightarrow X\times X be the natural transformation (33). Suppose that Φ\Phi respects the filtrations, i.e. for each ss, the natural transformation es(FsX)FsX×FsX\mathrm{es}(F_{s}X)\Rightarrow F_{s}X\times F_{s}X factors as

es(FsX)Fs(X×X)FsX×FsX.\mathrm{es}(F_{s}X)\Rightarrow F_{s}(X\times X)\Rightarrow F_{s}X\times F_{s}X.

Then the diagram

|X|\textstyle{|X|\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}|Φ|\scriptstyle{|\Phi|}|Φ|\scriptstyle{|\Phi|}|X|×|X|\textstyle{|X|\!\times\!|X|\ignorespaces\ignorespaces\ignorespaces\ignorespaces}|Φ|×id\scriptstyle{|\Phi|\times\mathrm{id}}|X|×|X|\textstyle{|X|\!\times\!|X|\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id×|Φ|\scriptstyle{\mathrm{id}\times|\Phi|}|X|×|X|×|X|\textstyle{|X|\!\times\!|X|\!\times\!|X|}

commutes up to a homotopy

[0,1]×|X||X|×|X|×|X|.[0,1]\times|X|\to|X|\times|X|\times|X|.

Moreover, the homotopy may be chosen so that, for each s0s\geq 0, it restricts to a map

[0,1]×|FsX||Fs(X×X×X)|.[0,1]\times|F_{s}X|\to|F_{s}(X\times X\times X)|.
Proof.

We shall construct a map

GX:[0,1]×|X||X|,G_{X}\colon[0,1]\times|X|\to|X|,

for every bisimplicial set XX, which is natural in XX. The map GXG_{X} will have the property that GX(0,):|X||X|G_{X}(0,-)\colon|X|\to|X| is the identity map, and, moreover, the composition

(37) |X|×[0,1]GX|X|(id×Φ)Φ|X|×|X|×|X||X|\times[0,1]\xrightarrow{G_{X}}|X|\xrightarrow{(\mathrm{id}\times\Phi)\circ\Phi}|X|\times|X|\times|X|

is a natural homotopy from (id×Φ)Φ(\mathrm{id}\times\Phi)\circ\Phi to (Φ×id)Φ(\Phi\times\mathrm{id})\circ\Phi.

Suppose we construct such maps GXG_{X}, natural in bisimplicial sets XX. We now explain how this is enough to prove the Proposition. Suppose that XX is a filtered bisimplicial set and suppose Φ:es(X)X×X\Phi\colon\mathrm{es}(X)\Rightarrow X\times X respects the filtration. Then for each s0s\geq 0, the map Φ\Phi restricts to a map

|Fs(X)||Fs(X×X)|,|F_{s}(X)|\to|F_{s}(X\times X)|,

and similarly, the map (id×Φ)Φ(\mathrm{id}\times\Phi)\circ\Phi, and also the map (Φ×id)Φ(\Phi\times\mathrm{id})\circ\Phi, restricts to a map

(38) |Fs(X)||Fs(X×X×X)|.|F_{s}(X)|\to|F_{s}(X\times X\times X)|.

for each s0s\geq 0. Then applying (37) to Fs(X)F_{s}(X) and combining with (38) yields a homotopy

|Fs(X)|×[0,1]|Fs(X×X×X)||F_{s}(X)|\times[0,1]\longrightarrow|F_{s}(X\times X\times X)|

from (id×Φ)Φ(\mathrm{id}\times\Phi)\circ\Phi to (Φ×id)Φ(\Phi\times\mathrm{id})\circ\Phi, as desired.

Now we construct the maps GXG_{X}. They will be defined first on the level of individual simplices, shown to be natural in face and degeneracy maps, and hence defined for all simplicial sets. Since they are natural in morphisms of simplicial sets, they then define maps on bisimplicial sets.

Consider Cartesian coordinates on the standard nn-simplex for each nn:

(39) Δn{(t1,,tn)n0t1t2tn1}.\Delta^{n}\cong\{(t_{1},\ldots,t_{n})\in\mathbb{R}^{n}\mid 0\leq t_{1}\leq t_{2}\leq\cdots\leq t_{n}\leq 1\}.

First let g:[0,1][0,1]g\colon[0,1]\to[0,1] be any nondecreasing continuous function with g(0)=0g(0)=0 and g(1)=1g(1)=1. Later on we will specialize gg to a particular piecewise linear function. By applying gg coordinatewise to (39), gg defines self-maps on each simplex gn:ΔnΔng_{n}\colon\Delta^{n}\to\Delta^{n} given by

(t1,,tn)(g(t1),,g(tn)).(t_{1},\ldots,t_{n})\mapsto(g(t_{1}),\ldots,g(t_{n})).

For each nn, the maps gng_{n} and gn1g_{n-1} are compatible with the n+1n+1 face maps Δn1Δn\Delta^{n-1}\to\Delta^{n}, which send (t1,,tn1)(t_{1},\ldots,t_{n-1}) to

(0,t1,,tn1),(t1,,ti,ti,,tn1) for i=1,,n1, and (t1,,tn1,1)(0,t_{1},\ldots,t_{n-1}),\quad(t_{1},\ldots,t_{i},t_{i},\ldots,t_{n-1})\text{ for $i=1,\ldots,n-1$, and }(t_{1},\ldots,t_{n-1},1)

respectively, as well as the nn degeneracy maps ΔnΔn1\Delta^{n}\to\Delta^{n-1} that send (t1,,tn)(t_{1},\ldots,t_{n}) to

(t1,,ti^,,tn)(t_{1},\ldots,\widehat{t_{i}},\ldots,t_{n})

for each i=1,,ni=1,\ldots,n, respectively. Therefore, gg determines maps gX:|X||X|g_{X}\colon|X|\to|X| for any simplicial set XX, naturally in XX. If XX is a bisimplicial set, then gg also determines a map |Xp,||Xp,||X_{p,\bullet}|\to|X_{p,\bullet}| for each pp, and by gluing these, we get a map gX:|X||X|g_{X}\colon|X|\to|X| as well, which is natural in morphisms of bisimplicial sets. Finally, let

(40) GX:[0,1]×|X||X|G_{X}\colon[0,1]\times|X|\to|X|

be the straight line homotopy from the identity map to gXg_{X}; this is again natural in morphisms of bisimplicial sets.

Now we specialize to a particular gg. Let g:[0,1][0,1]g\colon[0,1]\to[0,1] be the increasing, piecewise linear map that sends the intervals

[0,14],[14,12],[12,1]\big{[}0,\tfrac{1}{4}\big{]},\quad\big{[}\tfrac{1}{4},\tfrac{1}{2}\big{]},\quad\big{[}\tfrac{1}{2},1\big{]}

linearly to the intervals

[0,12],[12,34],[34,1]\big{[}0,\tfrac{1}{2}\big{]},\quad\big{[}\tfrac{1}{2},\tfrac{3}{4}\big{]},\quad\big{[}\tfrac{3}{4},1\big{]}

respectively. What remains only to show is that the composition

|X|gX|X|(id×Φ)Φ|X|×|X|×|X||X|\xrightarrow{g_{X}}|X|\xrightarrow{(\textrm{id}\times\Phi)\circ\Phi}|X|\times|X|\times|X|

is equal to (Φ×id)Φ.(\Phi\times\textrm{id})\circ\Phi. It suffices to perform this calculation in the case of a simplex

X=Δn=HomΔ(,[n]):Δop𝖲𝖾𝗍.X=\Delta^{n}=\operatorname{Hom}_{\Delta}(-,[n])\colon\Delta^{\mathrm{op}}\to\mathsf{Set}.

We first state the following lemma, without proof:

Lemma 3.12.

Let XX be an nn-simplex. In Cartesian coordinates, the map

Φ:|X||X|×|X|\Phi\colon|X|\to|X|\times|X|

is given by

(t1,,tn)((t1,,ti,1,,1),(0,,0,ti+1,,tn))(t_{1},\ldots,t_{n})\mapsto((t_{1}^{\prime},\ldots,t_{i}^{\prime},1,\ldots,1),(0,\ldots,0,t_{i+1}^{\prime},\ldots,t_{n}^{\prime}))

for

0t1ti12ti+1tn10\leq t_{1}\leq\cdots\leq t_{i}\leq\tfrac{1}{2}\leq t_{i+1}\leq\cdots\leq t_{n}\leq 1

and where tt^{\prime}_{\ell} denotes the result of applying to tt_{\ell} the appropriate linear rescaling

[0,12][0,1] or [12,1][0,1].\big{[}0,\tfrac{1}{2}\big{]}\to[0,1]\quad\text{ or }\quad\big{[}\tfrac{1}{2},1\big{]}\to[0,1].

Continue to let XX be an nn-simplex. Applying Lemma 3.12, in Cartesian coordinates, the map (id×Φ)Φ(\mathrm{id}\times\Phi)\circ\Phi is given by

(t1,,tn)((t1,,ti,1,,1),(0,,0,ti+1,,tj,1,,1),(0,,0,tj+1,,tn))(t_{1},\ldots,t_{n})\mapsto((t_{1}^{\prime},\ldots,t_{i}^{\prime},1,\ldots,1),(0,\ldots,0,t_{i+1}^{\prime},\ldots,t_{j}^{\prime},1,\ldots,1),(0,\ldots,0,t_{j+1}^{\prime},\ldots,t_{n}^{\prime}))

for

0t1ti12ti+1tj34tj+1tn10\leq t_{1}\leq\cdots\leq t_{i}\leq\tfrac{1}{2}\leq t_{i+1}\leq\cdots\leq t_{j}\leq\tfrac{3}{4}\leq t_{j+1}\leq\cdots\leq t_{n}\leq 1

and tt_{\ell}^{\prime} denotes the result of applying to tt_{\ell} the appropriate linear rescaling

[0,12][0,1],[12,34][0,1], or[34,1][0,1].\big{[}0,\tfrac{1}{2}\big{]}\to[0,1],\quad\big{[}\tfrac{1}{2},\tfrac{3}{4}\big{]}\to[0,1],\quad\text{ or}\quad\big{[}\tfrac{3}{4},1\big{]}\to[0,1].

And the map (Φ×id)Φ(\Phi\times\mathrm{id})\circ\Phi has the same description as above, replacing the numbers 12\tfrac{1}{2} and 34\tfrac{3}{4} with 14\tfrac{1}{4} and 12\tfrac{1}{2} respectively. The equality

(id×Φ)ΦgX=(Φ×id)Φ(\mathrm{id}\times\Phi)\circ\Phi\circ g_{X}=(\Phi\times\mathrm{id})\circ\Phi

is now clear. This proves Proposition 3.11. ∎

For X=NS(Proj)X=N_{\bullet}S_{\bullet}(\mathrm{Proj}_{\mathbb{Z}}) with Proj\mathrm{Proj}_{\mathbb{Z}} the Waldhausen category of finitely generated projective \mathbb{Z}-modules, |X|=BK()|X|=BK(\mathbb{Z}) and we will use the filtered map |X||es(X)||X|×|X||X|\approx|\mathrm{es}(X)|\to|X|\times|X| to induce a coproduct on the Quillen spectral sequence. The product will be induced by a filtered map |X|×|X||X||X|\times|X|\to|X| which is in some ways easier to comprehend, but depends on extra structure on the category Proj\mathrm{Proj}_{\mathbb{Z}}, namely the symmetric monoidal structure given by direct sum of free \mathbb{Z}-modules. Strictly speaking, the direct sum operation (P,P)PP(P,P^{\prime})\mapsto P\oplus P^{\prime} is a choice (at least of the set underlying PPP\oplus P^{\prime}), because of this, the definition of the product map |X|×|X||X||X|\times|X|\to|X| looks a bit lengthy when spelled out, but hopefully will not be surprising.

Proposition 3.13.

Let

X=NS(Proj):Δop×Δop𝖲𝖾𝗍X=N_{\bullet}S_{\bullet}(\mathrm{Proj}_{\mathbb{Z}})\colon\Delta^{\mathrm{op}}\times\Delta^{\mathrm{op}}\to\mathsf{Set}

be the bisimplicial set with

Xp,q=NqSp(Proj)X_{p,q}=N_{q}S_{p}(\mathrm{Proj}_{\mathbb{Z}})

as before. Define a product

(41) m:X×XXm\colon X\times X\to X

given by maps

(X×X)p,q=Xp,q×Xp,qXp,q(X\times X)_{p,q}=X_{p,q}\times X_{p,q}\to X_{p,q}

which are direct sums of composable chains of morphisms of triangular diagrams. Precisely, we first choose for each pair of objects V,VV,V^{\prime} of Proj\mathrm{Proj}_{\mathbb{Z}} an object m(V,V)m(V,V^{\prime}) and morphisms

Vm(V,V)VV\to m(V,V^{\prime})\leftarrow V^{\prime}

satisfying the universal property of coproducts. In other words, m(V,V)m(V,V^{\prime}) is a chosen model for the direct sum VVV\oplus V^{\prime}, and there is a canonical associator and symmetry making (Proj,m)(\mathrm{Proj}_{\mathbb{Z}},m) into a symmetric monoidal category. As usual, any two choices of mm’s will be canonically isomorphic, although the functions N0Proj×N0ProjN0ProjN_{0}\mathrm{Proj}_{\mathbb{Z}}\times N_{0}\mathrm{Proj}_{\mathbb{Z}}\to N_{0}\mathrm{Proj}_{\mathbb{Z}} need not be equal. Relatedly, the product map m:|X|×|X||X|m\colon|X|\times|X|\to|X| that we will define depends on the choice of mm, although its homotopy class as a filtered map will not. From now on we will fix such a choice and write simply VVV\oplus V^{\prime} for the chosen object m(V,V)m(V,V^{\prime}).

Suppose now we are given two elements of Xp,qX_{p,q}, given (after suppressing chosen cokernels from the notation, in other words suppressing from the notation all but the first row of each triangular diagram (14)) by rectangular diagrams

0V10Vp00V1qVpq,0W10Wp00W1qWpq.\lx@xy@svg{\hbox{\raise 2.5pt\hbox{\kern 5.5pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&&&\\&&&\\&&&\crcr}}}\ignorespaces{\hbox{\kern-5.5pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 22.57164pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 22.57164pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{V_{1}^{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 41.11333pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 58.18497pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 20.59248pt\raise-13.69083pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-1.75pt\hbox{$\scriptstyle{\cong}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 31.84248pt\raise-19.88165pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 58.18497pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 71.68497pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 88.7566pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 88.7566pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{V_{p}^{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 86.84555pt\raise-13.69083pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-1.75pt\hbox{$\scriptstyle{\cong}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 98.09555pt\raise-19.88165pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern-3.0pt\raise-27.38165pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{}$}}}}}}}{\hbox{\kern 25.09248pt\raise-27.38165pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{\vdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 20.59248pt\raise-40.78746pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-1.75pt\hbox{$\scriptstyle{\cong}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 31.84248pt\raise-46.85997pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 58.18497pt\raise-27.38165pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{\cdots}$}}}}}}}{\hbox{\kern 91.34555pt\raise-27.38165pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{\vdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 86.84555pt\raise-40.78746pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-1.75pt\hbox{$\scriptstyle{\cong}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 98.09555pt\raise-46.85997pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern-5.5pt\raise-54.19328pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 22.57164pt\raise-54.19328pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 22.57164pt\raise-54.19328pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{V_{1}^{q}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 41.11333pt\raise-54.19328pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 58.18497pt\raise-54.19328pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 58.18497pt\raise-54.19328pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 71.68497pt\raise-54.19328pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 88.7566pt\raise-54.19328pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 88.7566pt\raise-54.19328pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{V_{p}^{q}}$}}}}}}}\ignorespaces}}}}\ignorespaces,\qquad\lx@xy@svg{\hbox{\raise 2.5pt\hbox{\kern 5.5pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&&&\\&&&\\&&&\crcr}}}\ignorespaces{\hbox{\kern-5.5pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 22.57164pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 22.57164pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{W_{1}^{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 43.89111pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 60.96274pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 21.98137pt\raise-13.69083pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-1.75pt\hbox{$\scriptstyle{\cong}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 33.23137pt\raise-19.88165pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 60.96274pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 74.46274pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 92.92326pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 92.92326pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{W_{p}^{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 92.40111pt\raise-13.69083pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-1.75pt\hbox{$\scriptstyle{\cong}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 103.65111pt\raise-19.88165pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern-3.0pt\raise-27.38165pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{}$}}}}}}}{\hbox{\kern 26.48137pt\raise-27.38165pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{\vdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 21.98137pt\raise-40.78746pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-1.75pt\hbox{$\scriptstyle{\cong}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 33.23137pt\raise-46.85997pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 60.96274pt\raise-27.38165pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{\cdots}$}}}}}}}{\hbox{\kern 96.90111pt\raise-27.38165pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{\vdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 92.40111pt\raise-40.78746pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-1.75pt\hbox{$\scriptstyle{\cong}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 103.65111pt\raise-46.85997pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern-5.5pt\raise-54.19328pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 22.57164pt\raise-54.19328pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 22.57164pt\raise-54.19328pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{W_{1}^{q}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 43.89111pt\raise-54.19328pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 60.96274pt\raise-54.19328pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 60.96274pt\raise-54.19328pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 74.46274pt\raise-54.19328pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 91.53438pt\raise-54.19328pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 91.53438pt\raise-54.19328pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{W_{p}^{q}.}$}}}}}}}\ignorespaces}}}}\ignorespaces

The notation that has been suppressed includes a choice of cokernel Vi,jkV^{k}_{i,j} for each map VikVjkV^{k}_{i}\to V^{k}_{j}. For ease of notation, we shall abbreviate rectangular diagrams of this form further, so that the diagrams above are abbreviated as

0V1Vp,0W1Wp.0\to V_{1}^{\bullet}\to\cdots\to V_{p}^{\bullet},\qquad 0\to W_{1}^{\bullet}\to\cdots\to W_{p}^{\bullet}.

Then their product is defined to be the diagram

V1W1VpWp.V_{1}^{\bullet}\oplus W_{1}^{\bullet}\to\cdots\to V_{p}^{\bullet}\oplus W_{p}^{\bullet}.

The chosen cokernels, which still do not appear in the notation, are the direct sums of the existing choices: namely, the chosen cokernel for

VikWikVjkWjkV_{i}^{k}\oplus W_{i}^{k}\to V_{j}^{k}\oplus W_{j}^{k}

is Vi,jkWi,jkV_{i,j}^{k}\oplus W_{i,j}^{k}. Write m:|X|×|X||X|m\colon|X|\times|X|\to|X| and Φ:|X||X|×|X|\Phi\colon|X|\to|X|\times|X| also for the maps on spaces. Then we claim that product and coproduct are compatible:

|X|×|X|\textstyle{|X|\times|X|\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}m\scriptstyle{m}Φ×Φ\scriptstyle{\Phi\times\Phi}|X|\textstyle{|X|\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Φ\scriptstyle{\Phi}|X|×|X|×|X|×|X|\textstyle{|X|\!\times\!|X|\!\times\!|X|\!\times\!|X|\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(m×m)(id×s×id)\scriptstyle{(m\times m)\circ(\mathrm{id}\times s\times\mathrm{id})}|X|×|X|\textstyle{|X|\!\times\!|X|}

where s:X×XX×Xs\colon X\times X\to X\times X exchanges first and second coordinate. The lower horizontal map is thus

(x1,x2,x3,x4)(m(x1,x3),m(x2,x4)).(x_{1},x_{2},x_{3},x_{4})\mapsto(m(x_{1},x_{3}),m(x_{2},x_{4})).
Proof.

It suffices to show that the compositions

es(X)×es(X)𝑚es(X)ΦX×X\mathrm{es}(X)\times\mathrm{es}(X)\xrightarrow{m}\mathrm{es}(X)\xrightarrow{\Phi}X\times X

and

es(X)×es(X)Φ×ΦX×X×X×X(m×m)(id×s×id)X×X\mathrm{es}(X)\times\mathrm{es}(X)\xrightarrow{\Phi\times\Phi}X\times X\times X\times X\xrightarrow{(m\times m)\circ(\mathrm{id}\times s\times\mathrm{id})}X\times X

are equal as morphisms of bisimplicial sets. This is direct from the definitions: both maps send a pair of (p,q)(p,q)-simplices of es(X)\mathrm{es}(X)

((0V1V2p+1),(0W1W2p+1))((0\to V_{1}^{\bullet}\to\cdots\to V_{2p+1}^{\bullet}),(0\to W_{1}^{\bullet}\to\cdots\to W_{2p+1}^{\bullet}))

to

((0V1W1V2p+1W2p+1),(0Vp+1,p+2Wp+1,p+2Vp+1,2p+1Wp+1,2p+1)),\begin{aligned} ((0\to V_{1}^{\bullet}\oplus W_{1}^{\bullet}\to\cdots\to V_{2p+1}^{\bullet}\oplus W_{2p+1}^{\bullet}),(0\to V_{p+1,p+2}^{\bullet}\oplus W_{p+1,p+2}^{\bullet}\to\cdots\to V_{p+1,2p+1}^{\bullet}\oplus W_{p+1,2p+1}^{\bullet})),\end{aligned}

and the proposition follows. ∎

Remark 3.14.

We also record that the product mm is evidently compatible with the filtration on XX. Precisely, mm restricts to a map

(42) m:|FsX|×|FtX||Fs+tX|m\colon|F_{s}X|\times|F_{t}X|\to|F_{s+t}X|

for all s,t0s,t\geq 0.

Remark 3.15.

We also need the following generalities on multiplicative spectral sequences.

Given a filtered space and a product that respects filtrations, we obtain (at this level of generality) a spectral sequence on relative singular homology with products on each page, and all differentials on each page are derivations of the product (Leibniz rule), and the isomorphism from H(Er)Er+1H_{*}(E_{r})\to E_{r+1} is a multiplicative isomorphism with respect to the product on H(Er)H_{*}(E_{r}) induced from that on ErE_{r}.

Proposition 3.16.

Let

X=NS(Proj):Δop×Δop𝖲𝖾𝗍,X=N_{\bullet}S_{\bullet}(\mathrm{Proj}_{\mathbb{Z}})\colon\Delta^{\mathrm{op}}\times\Delta^{\mathrm{op}}\to\mathsf{Set},

with its rank filtration. Then the product m:X×XXm\colon X\times X\to X is commutative up to homotopy, respecting the filtration. In other words, there is a homotopy

H:|X|×|X|×[0,1]|X|H\colon|X|\times|X|\times[0,1]\to|X|

from mm to msm\circ s, where s:X×XX×Xs\colon X\times X\to X\times X switches coordinates. Furthermore, HH respects the filtration, in that it restricts to a map

H:|FtX|×|FuX|×[0,1]|Ft+uX|H\colon|F_{t}X|\times|F_{u}X|\times[0,1]\to|F_{t+u}X|

for all t,u0t,u\geq 0.

Proof.

We shall construct a morphism of bisimplicial sets

X×X×Δ1,0XX\times X\times\Delta^{1,0}\to X

whose geometric realization is the desired HH. Here

Δ1,0=HomΔ×Δ(,([1],[0])),\Delta^{1,0}=\operatorname{Hom}_{\Delta\times\Delta}(-,([1],[0])),

whose geometric realization is an interval [0,1][0,1]. For ease of notation, abbreviate

0V1Vp0\to V_{1}^{\bullet}\to\cdots\to V_{p}^{\bullet}

for a (p,q)(p,q)-simplex of XX, as before. The notation stands for a chain of qq morphisms of triangular diagrams of projective \mathbb{Z}-modules of size pp.

For i=0,,p+1i=0,\ldots,p+1, let fiΔ([p],[1])f_{i}\in\Delta([p],[1]) be the morphism with fi(x)=0f_{i}(x)=0 if x<ix<i and fi(x)=1f_{i}(x)=1 if xix\geq i. An element of (X×X×Δ1,0)p,q(X\times X\times\Delta^{1,0})_{p,q} is a triple

(0V1Vp,0W1Wp,fi)(0\to V_{1}^{\bullet}\to\cdots\to V_{p}^{\bullet},0\to W_{1}^{\bullet}\to\cdots\to W_{p}^{\bullet},f_{i})

for some i{0,p+1}i\in\{0,\ldots p+1\}. It is sent by HH to

0V1W1Vi1Wi1WiViWpVp.0\to V_{1}^{\bullet}\oplus W_{1}^{\bullet}\to\cdots\to V_{i-1}^{\bullet}\oplus W_{i-1}^{\bullet}\to W_{i}^{\bullet}\oplus V_{i}^{\bullet}\to\cdots\to W_{p}^{\bullet}\oplus V_{p}^{\bullet}.

These maps

(X×X×Δ1,0)p,qXp,q(X\times X\times\Delta^{1,0})_{p,q}\to X_{p,q}

are compatible with face and degeneracy maps, so we have produced a map of bisimplicial sets, which restricts to mm and msm\circ s at t=0t=0 and t=1t=1, respectively. Finally, HH obviously respects filtration: it restricts to a map of bisimplicial sets, for any t,u0t,u\geq 0,

FtX×FuX×Δ1,0Ft+uX,F_{t}X\times F_{u}X\times\Delta^{1,0}\to F_{t+u}X,

since the rank of a direct sum of modules of rank tt and uu is t+ut+u. ∎

3.2. Monoidality for the spectral sequence of a filtered space

We recall a certain monoidality of the spectral sequence associated to a filtered space. This seems well known in the algebraic topology literature, so we list the precise statements we need and give some references: [CE99, Dou59a, Dou59b]. See also [Rog, Chapter 6] and [Goe].

3.2.1. Filtered spaces

In this section we write HH_{*} for singular homology with coefficients in some field. Recall that for a topological space XX filtered by subcomplexes FtXXF_{t}X\subset X, there is a spectral sequence

Es,t1=Es,t1(X)=Hs+t(FtX,Ft1X)H(X),E^{1}_{s,t}=E^{1}_{s,t}(X)=H_{s+t}(F_{t}X,F_{t-1}X)\Rightarrow H_{*}(X),

with convergence assuming the filtration is bounded below, for instance F1X=F_{-1}X=\emptyset, and exhaustive: colimtH(FtX)H(X)\operatorname*{colim}_{t}H_{*}(F_{t}X)\to H_{*}(X) is an isomorphism. For definiteness, let us work with the construction of this spectral sequence given in [CE99, Chap. XV, §7], see especially Example 3 on page 335. This agrees with the construction in [Dou59a, Section II.C] apart from notation (in particular, the latter writes π(p,q)\pi(p,q) for what the former denotes H(p,q)H(p,q)). This spectral sequence is natural with respect to all maps of filtered spaces: that is, continuous maps XXX\to X^{\prime} sending FtXF_{t}X into FtXF_{t}X^{\prime}.

3.2.2. Products of filtered spaces

If XX^{\prime} and X′′X^{\prime\prime} are filtered spaces, then the Cartesian product X=X×X′′X=X^{\prime}\times X^{\prime\prime} inherits a filtration, namely

FtX=Im(u+vtFuX×FvX′′X×X′′=X).F_{t}X=\mathrm{Im}\Big{(}\coprod_{u+v\leq t}F_{u}X^{\prime}\times F_{v}X^{\prime\prime}\to X^{\prime}\times X^{\prime\prime}=X\Big{)}.

In this situation the inclusion FuX×FvX′′Fu+vXF_{u}X^{\prime}\times F_{v}X^{\prime\prime}\hookrightarrow F_{u+v}X induces a chain map

C(FuX)C(FvX′′)C(Fu+vX)C_{*}(F_{u}X^{\prime})\otimes C_{*}(F_{v}X^{\prime\prime})\to C_{*}(F_{u+v}X)

defined by the chain-level cross product. This inclusion sends both subspaces FuX×Fv1X′′F_{u}X^{\prime}\times F_{v-1}X^{\prime\prime} and Fu1X×FvX′′F_{u-1}X^{\prime}\times F_{v}X^{\prime\prime} into Fu+v1XF_{u+v-1}X, so the cross product factors over a chain map

(43) C(FuX,Fu1X)C(FvX′′,Fv1X′′)C(Fu+vX,Fu+v1X).C_{*}(F_{u}X^{\prime},F_{u-1}X^{\prime})\otimes C_{*}(F_{v}X^{\prime\prime},F_{v-1}X^{\prime\prime})\to C_{*}(F_{u+v}X,F_{u+v-1}X).

Passing to homology then gives a homomorphism

(44) ϕ1:Ep,q1(X)Ep,q1(X′′)Ep+p,q+q1(X)\phi^{1}\colon E^{1}_{p,q}(X^{\prime})\otimes E^{1}_{p^{\prime},q^{\prime}}(X^{\prime\prime})\to E^{1}_{p+p^{\prime},q+q^{\prime}}(X)

which we will call the exterior product. In brief, these make XE,1(X)X\mapsto E^{1}_{*,*}(X) into a lax monoidal functor from filtered spaces to bigraded vector spaces.

Similarly, filtering H(X)H_{*}(X) by the images of the H(FtX)H_{*}(F_{t}X), and similarly for H(X)H_{*}(X^{\prime}) and H(X′′)H_{*}(X^{\prime\prime}), the cross product H(X)H(X′′)H(X)H_{*}(X^{\prime})\otimes H_{*}(X^{\prime\prime})\to H_{*}(X) descends to a pairing

(45) GrpHp+q(X)GrpHp+q(X′′)Grp+pHp+p+q+q(X).\mathrm{Gr}_{p}H_{p+q}(X^{\prime})\otimes\mathrm{Gr}_{p^{\prime}}H_{p^{\prime}+q^{\prime}}(X^{\prime\prime})\to\mathrm{Gr}_{p+p^{\prime}}H_{p+p^{\prime}+q+q^{\prime}}(X).
Proposition 3.17.

In the above setting there are pairings

ϕr:Ep,qr(X)Ep,qr(X′′)Ep+p,q+qr(X)\phi^{r}\colon E^{r}_{p,q}(X^{\prime})\otimes E^{r}_{p^{\prime},q^{\prime}}(X^{\prime\prime})\to E^{r}_{p+p^{\prime},q+q^{\prime}}(X)

for all r1r\geq 1, agreeing with (44) for r=1r=1, satisfying the Leibniz rule

drϕr(xx′′)=ϕr((drx)x′′+(1)p+qx(drx′′))d^{r}\phi^{r}(x^{\prime}\otimes x^{\prime\prime})=\phi^{r}((d^{r}x^{\prime})\otimes x^{\prime\prime}+(-1)^{p+q}x^{\prime}\otimes(d^{r}x^{\prime\prime}))

and such that the induced pairing on homology of rrth pages is identified with ϕr+1\phi^{r+1}. (Such pairings are of course unique if they exist since each determines the next—the main content is the Leibniz rule so that ϕr\phi^{r} descends to a pairing on homology of rrth pages.)

Moreover, the induced pairing ϕ\phi^{\infty} of EE^{\infty}-pages is identified with (45). Finally, if XX^{\prime} and X′′X^{\prime\prime} are CW complexes filtered by subcomplexes, then the homomorphisms

ϕr:E,r(X)E,r(X′′)E,r(X),\phi^{r}\colon E^{r}_{*,*}(X^{\prime})\otimes E^{r}_{*,*}(X^{\prime\prime})\to E^{r}_{*,*}(X),

obtained by taking direct sum over all p,q,p,qp,q,p^{\prime},q^{\prime}, are isomorphisms for all r1r\geq 1.

Proof sketch.

This is mostly contained in [Dou59b, Théorème II.A]. In the notation from there, we should set

π(q,r)=nHn(FqX,FrX)\pi(q,r)=\bigoplus_{n}H_{n}(F_{-q}X,F_{-r}X)

for qr-\infty\leq q\leq r\leq\infty (where we write FX=F_{-\infty}X=\emptyset and FX=XF_{\infty}X=X), let

η:π(q,r)π(q,r)\eta\colon\pi(q,r)\to\pi(q^{\prime},r^{\prime})

be defined by functoriality of homology for qqq^{\prime}\leq q and rrr^{\prime}\leq r, and let

:π(q,r)π(r,s)\partial\colon\pi(q,r)\to\pi(r,s)

be the connecting homomorphism for the triple (FqX,FrX,FsX)(F_{q}X,F_{r}X,F_{s}X) for qrs-\infty\leq q\leq r\leq s\leq\infty. This data is called a système spectraux in [Dou59a] and a Cartan–Eilenberg system in many other places, and satisfies the axioms (SP.1)–(SP.5) of [CE99].

Defining systèmes spectraux π\pi^{\prime} and π′′\pi^{\prime\prime} associated to the filtered spaces XX^{\prime} and X′′X^{\prime\prime} in the same manner, the inclusions FnX×FqX′′FnqXF_{-n}X^{\prime}\times F_{-q}X^{\prime\prime}\hookrightarrow F_{-n-q}X then induce homomorphisms

π(n,n+r)π′′(q,q+r)ϕrπ(n+q,n+q+r)=H(FnqX,FnqrX)\pi^{\prime}(n,n+r)\otimes\pi^{\prime\prime}(q,q+r)\xrightarrow{\phi_{r}}\pi(n+q,n+q+r)=H_{*}(F_{-n-q}X,F_{-n-q-r}X)

for all r1r\geq 1, in the same way as (44) which is the special case r=1r=1. These homomorphisms satisfy the assumptions of Douady’s theorem, which then gives the stated result, apart from the claim that the ϕr\phi^{r} define isomorphisms after taking direct sum over all bidegrees. The original reference [Dou59b] in fact omits the proof, but details can be found elsewhere, for instance [Goe] or [Rog]. The verification of axioms (SPP.1) and (SPP.2) in Douady’s theorem is as in [Rog, Proposition 6.3.12].

It remains to see that the exterior products ϕr\phi^{r} become isomorphisms after taking direct sum over all bidegrees, when XX^{\prime} and X′′X^{\prime\prime} are CW complexes filtered by subcomplexes. In that case we can identify relative homology with reduced homology of the quotient space: Ep,q1(X)=H~p+q(FpX/Fp1X)E^{1}_{p,q}(X^{\prime})=\widetilde{H}_{p+q}(F_{p}X^{\prime}/F_{p-1}X^{\prime}), and similarly for X′′X^{\prime\prime} and XX. Then ϕ1\phi^{1} is induced by maps (FuX/Fu1X)(FvX′′/Fv1X′′)Fu+vX/Fu+v1X(F_{u}X^{\prime}/F_{u-1}X^{\prime})\wedge(F_{v}X^{\prime\prime}/F_{v-1}X^{\prime\prime})\to F_{u+v}X/F_{u+v-1}X arising from the inclusions FuX×FvX′′Fu+vXF_{u}X^{\prime}\times F_{v}X^{\prime\prime}\hookrightarrow F_{u+v}X. Taking wedge sum over all filtrations, we see that ϕ1\phi^{1} is induced from a single map of pointed spaces

(sFsX/Fs1X)(sFsX′′/Fs1X′′)sFsX/Fs1X,\Big{(}\bigvee_{s\in\mathbb{Z}}F_{s}X^{\prime}/F_{s-1}X^{\prime}\Big{)}\wedge\Big{(}\bigvee_{s\in\mathbb{Z}}F_{s}X^{\prime\prime}/F_{s-1}X^{\prime\prime}\Big{)}\longrightarrow\bigvee_{s\in\mathbb{Z}}F_{s}X/F_{s-1}X,

and the claim about isomorphism follows for r=1r=1 from the Künneth theorem and the observation that this map is in fact a homeomorphism when X=X×X′′X=X^{\prime}\times X^{\prime\prime} is given the CW topology. But then inductively ϕr\phi^{r} is also an isomorphism for higher rr, again by the Künneth theorem. ∎

3.3. Hopf algebra structure on the Quillen spectral sequence

The results earlier in this section combine to yield the following extra structures on EQ{}^{Q}\!E, the homological Quillen spectral sequence.

Theorem 3.18.

  1. (1)

    The homological Quillen spectral sequence

    Es,t1Q=Ht(GLs(),Sts)H(BK()){}^{Q}\!E^{1}_{s,t}=H_{t}(\mathrm{GL}_{s}(\mathbb{Z}),\mathrm{St}_{s}\otimes\mathbb{Q})\Rightarrow H_{\ast}(BK(\mathbb{Z}))

    is a spectral sequence of Hopf algebras. That is, for each r0r\geq 0, there are maps

    mr:Es,trEs,trEs+s,t+tr,Δr:Es,trs+s′′=st+t′′=tEs,trEs′′,t′′r,m^{r}\colon E^{r}_{s,t}\otimes E^{r}_{s^{\prime},t^{\prime}}\to E^{r}_{s+s^{\prime},t+t^{\prime}},\qquad\Delta^{r}\colon E^{r}_{s,t}\to\bigoplus_{\begin{subarray}{c}s^{\prime}+s^{\prime\prime}=s\\ t^{\prime}+t^{\prime\prime}=t\end{subarray}}E^{r}_{s^{\prime},t^{\prime}}\otimes E^{r}_{s^{\prime\prime},t^{\prime\prime}},

    induced from

    m:BK()×BK()BK()andΦ:BK()BK()×BK()m\colon BK(\mathbb{Z})\times BK(\mathbb{Z})\to BK(\mathbb{Z})\quad\text{and}\quad\Phi\colon BK(\mathbb{Z})\to BK(\mathbb{Z})\times BK(\mathbb{Z})

    respectively, making ErE^{r} a bigraded Hopf algebra. Moreover, the differential

    dr:Es,trEsr,t+r1rd^{r}\colon E^{r}_{s,t}\to E^{r}_{s-r,t+r-1}

    is compatible with product and coproduct in the sense that

    dr(x1x2)=dr(x1)x2+(1)p1+q1xdr(x2)d^{r}(x_{1}\cdot x_{2})=d^{r}(x_{1})\cdot x_{2}+(-1)^{p_{1}+q_{1}}x\cdot d^{r}(x_{2})

    for xiEpi,qirx_{i}\in E^{r}_{p_{i},q_{i}}, where xyx\cdot y denotes the product mrm^{r}; and

    Δrdr=(dr1±1dr)Δr.\Delta^{r}\circ d^{r}=(d^{r}\otimes 1\pm 1\otimes d^{r})\circ\Delta^{r}.

    More precisely, the sign “±\pm” can be expressed in terms of the symmetry

    T:Ep,qrEp,qr\displaystyle T\colon E^{r}_{p,q}\otimes E^{r}_{p^{\prime},q^{\prime}} Ep,qrEp,qr\displaystyle\xrightarrow{\cong}E^{r}_{p^{\prime},q^{\prime}}\otimes E^{r}_{p,q}
    xy\displaystyle x\otimes y\quad (1)(p+q)(p+q)yx;\displaystyle\mapsto(-1)^{(p+q)(p^{\prime}+q^{\prime})}y\otimes x;

    the Leibniz rule for the coproduct should then read Δrdr=(dr1+T(dr1)T)Δr\Delta^{r}\circ d^{r}=(d^{r}\otimes 1+T\circ(d^{r}\otimes 1)\circ T)\circ\Delta^{r}.

  2. (2)

    There are isomorphisms E0,0r\mathbb{Q}\to E^{r}_{0,0}\to\mathbb{Q} acting as unit and counit respectively, making each page ErE^{r} into a bigraded Hopf algebra.

  3. (3)

    Giving the homology of the rrth page the Hopf algebra structure induced by mrm^{r} and Δr\Delta^{r}, the isomorphism Er+1=H(Er,dr)E^{r+1}=H(E^{r},d^{r}) is a Hopf algebra isomorphism.

  4. (4)

    The filtration of H(BK();)H_{*}(BK(\mathbb{Z});\mathbb{Q}) induced by convergence of the spectral sequence is multiplicative and comultiplicative, where the product on H(BK();)H_{*}(BK(\mathbb{Z});\mathbb{Q}) is induced by m:BK()×BK()BK()m\colon BK(\mathbb{Z})\times BK(\mathbb{Z})\to BK(\mathbb{Z}) (in turn constructed from the \oplus operation on the Waldhausen construction; this also agrees with the product induced by the loop space structure) and the coproduct on H(BK();)H_{*}(BK(\mathbb{Z});\mathbb{Q}) is induced by the diagonal map (dual to cup product). With respect to the induced Hopf algebra structure on GrH(BK();)\mathrm{Gr}H_{*}(BK(\mathbb{Z});\mathbb{Q}), the isomorphisms Ep,q=GrpHp+q(BK();)E^{\infty}_{p,q}=\mathrm{Gr}_{p}H_{p+q}(BK(\mathbb{Z});\mathbb{Q}) form an isomorphism of bigraded Hopf algebras.

  5. (5)

    The product on ErE^{r} is graded-commutative for all r1r\geq 1:

    x1x2=(1)(p1+q1)(p2+q2)x2x1,x_{1}\cdot x_{2}=(-1)^{(p_{1}+q_{1})(p_{2}+q_{2})}x_{2}\cdot x_{1},

    for xix_{i} in Epi,qirE^{r}_{p_{i},q_{i}}.

  6. (6)

    The coproduct Δr\Delta^{r} is graded co-commutative for r=r=\infty but not necessarily for finite rr. (We are of course asserting that it is co-associative for all rr, which is part of the axioms for Hopf algebras.)

Proof.

Setting X=X′′=BK()X^{\prime}=X^{\prime\prime}=BK(\mathbb{Z}) in Proposition 3.17, we obtain isomorphisms

(46) E,r(BK())E,r(BK())ϕrE,r(BK()×BK())E^{r}_{*,*}(BK(\mathbb{Z}))\otimes E^{r}_{*,*}(BK(\mathbb{Z}))\xrightarrow{\phi^{r}}E^{r}_{*,*}(BK(\mathbb{Z})\times BK(\mathbb{Z}))

satisfying the Leibniz rule on each page, as well as compatibility between pages and with the abutments.

Now we combine with the space-level products and coproducts and functoriality of the spectral sequence with respect to filtered maps, using Corollary 3.9, Propositions 3.113.133.16, and Remark 3.14. For instance, the space-level coproduct Φ:BK()BK()×BK()\Phi\colon BK(\mathbb{Z})\to BK(\mathbb{Z})\times BK(\mathbb{Z}) is a map of filtered spaces and hence induces a map of spectral sequences

E,r(BK())ΦE,r(BK()×BK())E^{r}_{*,*}(BK(\mathbb{Z}))\xrightarrow{\Phi_{*}}E^{r}_{*,*}(BK(\mathbb{Z})\times BK(\mathbb{Z}))

which we combine with the inverse of (46) to obtain a coproduct on the rrth pages of the Quillen spectral sequence. Similarly for the product, while the unit and counit come from space-level maps {point}BK(){point}\{\text{point}\}\to BK(\mathbb{Z})\to\{\text{point}\}, where the one-point space is filtered as F1={point}=F0F_{-1}=\emptyset\subset\{\text{point}\}=F_{0}.

To see that the coproduct on the spectral sequence is coassociative, we write I=[0,1]I=[0,1] filtered as =F1IF0I=I\emptyset=F_{-1}I\subset F_{0}I=I and first observe that the two injections BK()I×BK()BK(\mathbb{Z})\hookrightarrow I\times BK(\mathbb{Z}), given by x(0,x)x\mapsto(0,x) and x(1,x)x\mapsto(1,x), induce equal maps of spectral sequences, since both are one-sided inverses to the isomorphism of spectral sequences induced by the projection I×BK()BK()I\times BK(\mathbb{Z})\to BK(\mathbb{Z}). Coassociativity then follows from the space-level homotopy

I×BK()BK()×BK()×BK(),I\times BK(\mathbb{Z})\to BK(\mathbb{Z})\times BK(\mathbb{Z})\times BK(\mathbb{Z}),

observing that this is in fact a filtered map.

To see that the coproduct is co-commutative on the EE^{\infty}-page, we use that the space-level map Φ\Phi is homotopic to the diagonal map. Therefore the induced coproduct Φ:H(BK())H(BK())H(BK())\Phi_{*}\colon H_{*}(BK(\mathbb{Z}))\to H_{*}(BK(\mathbb{Z}))\otimes H_{*}(BK(\mathbb{Z})) is co-commutative, but then this also holds for the induced map of associated gradeds, which is identified with ϕ:E,E,E,\phi^{\infty}\colon E^{\infty}_{*,*}\to E^{\infty}_{*,*}\otimes E^{\infty}_{*,*}.

All other properties follow from the corresponding space-level properties in a similar way. ∎

4. Proof of Theorems 1.1 and 1.2 and Corollary 1.9

4.1. A product on tropical moduli spaces

Towards a proof of Theorem 1.1, we shall first study a graded-commutative product on W0Hc(𝒜)W_{0}H^{*}_{c}(\mathcal{A}), which can be interpreted both in terms of products of abelian varieties and of tropical abelian varieties. As in the proof of Proposition 2.8, we have short exact sequences

(47) 0HkBM(Agtrop)𝜄HkBM(Agtrop,Ag1trop)Hk1BM(Ag1trop)00\to H^{\mathrm{BM}}_{k}(A_{g}^{\mathrm{trop}})\xrightarrow{\iota}H^{\mathrm{BM}}_{k}(A_{g}^{\mathrm{trop}},A_{g-1}^{\mathrm{trop}})\xrightarrow{\partial}H_{k-1}^{\mathrm{BM}}(A_{g-1}^{\mathrm{trop}})\to 0

for all kk and gg. These follow from the fact that HBM(Ag1trop)HBM(Agtrop)H_{*}^{\mathrm{BM}}(A_{g-1}^{\mathrm{trop}})\to H_{*}^{\mathrm{BM}}(A_{g}^{\mathrm{trop}}) is zero, established in (12).

Now let (P(g)[1],d)(P^{(g)}[-1],d) denote the degree-shifted perfect cone complex, whose definition and properties we shall now recall from [BBC+24]. The complex P(g)[1]P^{(g)}[-1] is a rational chain complex with differential of degree 1-1, with generators [σ,ω][\sigma,\omega] in degree dim(σ)\dim(\sigma), where σ\sigma is a perfect cone and ω\omega is an orientation of the linear span of σ\sigma. Relations are given by [σ,ω]=±[σ,ω][\sigma,\omega]=\pm[\sigma^{\prime},\omega^{\prime}] if σ\sigma and σ\sigma^{\prime} are in the same GLg()\mathrm{GL}_{g}(\mathbb{Z})-orbit, with a plus sign if the induced action of GLg()\mathrm{GL}_{g}(\mathbb{Z}) on the orientation ω\omega equals ω\omega^{\prime}, and a minus sign if not. The boundary of [σ,ω][\sigma,\omega] is a sum of codimension 11 faces of σ\sigma with induced orientation. Then from [BBC+24] we have HkBM(Agtrop)Hk(P(g)[1])H_{k}^{\mathrm{BM}}(A_{g}^{\mathrm{trop}})\cong H_{k}(P^{(g)}[-1]). Moreover, there are natural inclusions P(g1)P(g)P^{(g-1)}\to P^{(g)} such that

Hk(P(g1)[1])\textstyle{H_{k}(P^{(g-1)}[-1])\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}Hk(P(g)[1])\textstyle{H_{k}(P^{(g)}[-1])\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}HkBM(Ag1trop)\textstyle{H_{k}^{\mathrm{BM}}(A_{g-1}^{\mathrm{trop}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}HkBM(Agtrop)\textstyle{H_{k}^{\mathrm{BM}}(A_{g}^{\mathrm{trop}})}

commutes. Therefore from (47) we obtain short exact sequences

0Hk(P(g)[1])𝜄Hk((P(g)/P(g1))[1])Hk1(P(g1)[1])00\to H_{k}(P^{(g)}[-1])\xrightarrow{\iota}H_{k}((P^{(g)}/P^{(g-1)})[-1])\xrightarrow{\partial}H_{k-1}(P^{(g-1)}[-1])\to 0

for all kk and gg.

Now we construct a product map and prove a Leibniz rule. Consider the natural continuous maps

(48) Agtrop×AhtropAg+htropA_{g}^{\mathrm{trop}}\times A_{h}^{\mathrm{trop}}\to A_{g+h}^{\mathrm{trop}}

induced by block sum of positive semidefinite forms. These maps extend to

(Agtrop{},)(Ahtrop{},)(Ag+htrop{},),(A_{g}^{\mathrm{trop}}\cup\{\infty\},\infty)\wedge(A_{h}^{\mathrm{trop}}\cup\{\infty\},\infty)\to(A_{g+h}^{\mathrm{trop}}\cup\{\infty\},\infty),

the one-point compactifications with added point \infty. Hence we obtain

(49) HkBM(Agtrop)HBM(Ahtrop)Hk+BM(Ag+htrop).H_{k}^{\mathrm{BM}}(A_{g}^{\mathrm{trop}})\otimes H_{\ell}^{\mathrm{BM}}(A_{h}^{\mathrm{trop}})\to H_{k+\ell}^{\mathrm{BM}}(A_{g+h}^{\mathrm{trop}}).

Similarly, we have

(50) HkBM(Agtrop,Ag1trop)HBM(Ahtrop,Ah1trop)Hk+BM(Ag+htrop,Ag+h1trop)H_{k}^{\mathrm{BM}}(A_{g}^{\mathrm{trop}},A_{g-1}^{\mathrm{trop}})\otimes H_{\ell}^{\mathrm{BM}}(A_{h}^{\mathrm{trop}},A_{h-1}^{\mathrm{trop}})\to H_{k+\ell}^{\mathrm{BM}}(A_{g+h}^{\mathrm{trop}},A_{g+h-1}^{\mathrm{trop}})

Both (49) and (50) are induced from

(51) m:P(g)[1]P(g)[1]P(g+g)[1],[σ,ω][σ,ω][σ×σ,ωω]m\colon P^{(g)}[-1]\otimes P^{(g^{\prime})}[-1]\to P^{(g+g^{\prime})}[-1],\quad[\sigma,\omega]\otimes[\sigma^{\prime},\omega^{\prime}]\mapsto[\sigma\times\sigma^{\prime},\omega\ast\omega^{\prime}]

where, for cones σSym2((g))\sigma\subset\operatorname{Sym}^{2}((\mathbb{R}^{g})^{\vee}) and σSym2((g))\sigma^{\prime}\subset\operatorname{Sym}^{2}((\mathbb{R}^{g^{\prime}})^{\vee}), we take σ×σSym2((g))×Sym2((g))Sym2((g+g))\sigma\times\sigma^{\prime}\in\operatorname{Sym}^{2}((\mathbb{R}^{g^{\prime}})^{\vee})\times\operatorname{Sym}^{2}((\mathbb{R}^{g})^{\vee})\subset\operatorname{Sym}^{2}((\mathbb{R}^{g+g^{\prime}})^{\vee}) as the cone of block sums of symmetric bilinear forms in σ\sigma and σ\sigma^{\prime}. The product m(σ,σ)m(\sigma,\sigma^{\prime}) shall be denoted by σ.σ\sigma.\sigma^{\prime} henceforth, and similarly for the products (49) and (50). Note that mm is graded-commutative.

Consider the bigraded vector space 𝒜BM\mathcal{A}^{BM} defined as

𝒜s,tBM=Hs+tBM(Astrop;).\mathcal{A}^{BM}_{s,t}=H^{BM}_{s+t}(A^{\mathrm{trop}}_{s};\mathbb{Q}).

It is the (bi)graded dual of W0Hc(𝒜)W_{0}H^{*}_{c}(\mathcal{A}) and vanishes for t<0t<0. Then we have the following conclusion.

Proposition 4.1.

The maps (49) equip 𝒜BM\mathcal{A}^{\mathrm{BM}} with the structure of a \mathbb{Q}-algebra, which is graded-commutative with respect to its total grading.

Proof.

The graded-commutativity of the product on 𝒜BM\mathcal{A}^{\mathrm{BM}} follows from graded-commutativity of the product (51) on perfect cone complexes.∎

Remark 4.2.

The algebraic moduli space 𝒜\mathcal{A} has a natural space-level commutative product, given on connected components by

(52) 𝒜g×𝒜h𝒜g+h;(A1,A2)A1×A2.\mathcal{A}_{g}\times\mathcal{A}_{h}\to\mathcal{A}_{g+h};\ \ \ (A_{1},A_{2})\mapsto A_{1}\times A_{2}.

This product morphism 𝒜×𝒜𝒜\mathcal{A}\times\mathcal{A}\to\mathcal{A} is proper, and the induced pullback map

W0Hc(𝒜)W0Hc(𝒜)W0Hc(𝒜)W_{0}H^{*}_{c}(\mathcal{A})\to W_{0}H^{*}_{c}(\mathcal{A})\otimes W_{0}H^{*}_{c}(\mathcal{A})

agrees with (49) via the standard comparison isomorphisms and dualities.

To see this agreement, one may first note that the tropicalization of the algebraic product (52) is the tropical product defined in (48), induced by block sum of positive semidefinite quadratic forms. Then, use the fact that the Berkovich analytic skeleton of a product of abelian varieties over a valued field is the product of the skeletons, as principally polarized tropical abelian varieties. This is because the skeleton can be read off via non-archimedean analytic uniformization from the Raynaud cross diagram, as in [FRSS18, Section 3.2], and the Raynaud cross of the product is the product of the Raynaud crosses of the factors.

We also note that the algebraic product (52) extends to additive families of toroidal compactifications, such as the perfect cone compactifications. The fact that (48) is the tropicalization of (52) is well-known in this context. See [GHT18, Proposition 9].

The following version of the Leibniz rule will be used in the next subsection.

Proposition 4.3.

Let αHkBM(Agtrop,Ag1trop)\alpha\in H_{k}^{\mathrm{BM}}(A_{g}^{\mathrm{trop}},A_{g-1}^{\mathrm{trop}}) and βHBM(Ahtrop,Ah1trop)\beta\in H_{\ell}^{\mathrm{BM}}(A_{h}^{\mathrm{trop}},A_{h-1}^{\mathrm{trop}}). Then, with ι\iota and \partial as in (47), we have

ι(α.β)=(ια).β+(1)deg(α)α.(ιβ).\iota\partial(\alpha.\beta)=(\iota\partial\alpha).\beta+(-1)^{\mathrm{deg}(\alpha)}\alpha.(\iota\partial\beta).

Furthermore, the map ι\iota is a homomorphism for the multiplication maps (49) and (50).

Proof.

The key point is that the product (51) satisfies the graded Leibniz rule. More precisely, (51) is a morphism of chain complexes:

d([σ,ω].[σ,ω])=d[σ,ω].[σ,ω]+(1)dim(σ)[σ,ω].d[σ,ω]d([\sigma,\omega].[\sigma^{\prime},\omega^{\prime}])=d[\sigma,\omega].[\sigma^{\prime},\omega^{\prime}]+(-1)^{\mathrm{dim}(\sigma)}[\sigma,\omega].d[\sigma^{\prime},\omega^{\prime}]

by properties of faces of products of polyhedral cones. Then the Proposition follows since the product (50) is induced from this product map on perfect cone complexes. ∎

4.2. Hopf algebra structure on W0Hc(𝒜)W_{0}H^{*}_{c}(\mathcal{A})

Let E1Q{}^{Q}\!E^{1} denote the E1E^{1}-page of the homological Quillen spectral sequence. It is a graded-commutative bigraded Hopf algebra by Theorem 3.18. The graded-commutativity is with respect to the total degree.

The vector space E1,01Q{}^{Q}\!E^{1}_{1,0} is 1-dimensional and plays an important role in the Hopf algebra structure on E1Q{}^{Q}\!E^{1}, so we pause briefly to establish a canonical choice of generator. The fundamental group π1(BK())=K0()\pi_{1}(BK(\mathbb{Z}))=K_{0}(\mathbb{Z})\cong\mathbb{Z} is infinite cyclic, and is in fact generated by the loop corresponding to any 1-simplex (0P0,1)N0S1(Proj)(0\subset P_{0,1})\in N_{0}S_{1}(\mathrm{Proj}_{\mathbb{Z}}) for P0,1P_{0,1} of rank 1 (that is, P0,1P_{0,1} is isomorphic as an abelian group to \mathbb{Z}). Any two choices of P0,1P_{0,1} will be isomorphic, and lead to homotopic loops in the based space BK()BK(\mathbb{Z}). Let us write

e:S1BK()e\colon S^{1}\to BK(\mathbb{Z})

for any based loop in this homotopy class. The loop ee can also be interpreted as the generator of the fundamental group of gr1BK()=(F1BK())/(F0BK())\mathrm{gr}_{1}BK(\mathbb{Z})=(F_{1}BK(\mathbb{Z}))/(F_{0}BK(\mathbb{Z})). Hence we obtain a canonical generator

e([S1])H1(F1BK(),F0BK();)=E1,01Q,e_{*}([S^{1}])\in H_{1}(F_{1}BK(\mathbb{Z}),F_{0}BK(\mathbb{Z});\mathbb{Q})={}^{Q}\!E^{1}_{1,0}\cong\mathbb{Q},

where [S1][S^{1}] denotes the fundamental class of the circle S1=Δ1/(Δ1)S^{1}=\Delta^{1}/(\partial\Delta^{1}). This class is a permanent cycle in the spectral sequence, and we will henceforth use the notation eE1,01Qe\in{}^{Q}\!E^{1}_{1,0} for this homology class, as well as for its image eE1,0rQe\in{}^{Q}\!E^{r}_{1,0} in all subsequent pages rr\leq\infty.

Proposition 4.4.

With eE1,01Qe\in{}^{Q}\!E^{1}_{1,0} as above, there is a canonical isomorphism of graded-commutative algebras

𝒜BM[ϵ]/ϵ2E1Q,\mathcal{A}^{BM}\otimes_{\mathbb{Q}}\mathbb{Q}[\epsilon]/\epsilon^{2}\overset{\sim}{\to}{}^{Q}\!E^{1}\ ,

sending ϵ\epsilon to ee, where [ϵ]/ϵ2ϵ\mathbb{Q}[\epsilon]/\epsilon^{2}\cong\bigwedge\mathbb{Q}\epsilon is the graded exterior algebra generated by ϵ\epsilon.

Proof.

Let eE1,01Qe\in{}^{Q}\!E^{1}_{1,0}\cong\mathbb{Q} be the generator chosen above. (Any non-zero multiple thereof would work equally well for the following argument.) Since the coproduct Δ\Delta on E1Q{}^{Q}\!E^{1} respects the bigrading, we must have

(53) Δe=1e+e1.\Delta\,e=1\otimes e+e\otimes 1\ .

By graded-commutativity, or using the fact that E2,01Q{}^{Q}\!E^{1}_{2,0} vanishes, we have e2=0e^{2}=0.

Recall from (47) the short exact sequences

(54) 0Hs+tBM(Astrop)𝜄Hs+tBM(Astrop,As1trop)Hs+t1BM(As1trop)0,0\to H^{\mathrm{BM}}_{s+t}(A_{s}^{\mathrm{trop}})\xrightarrow{\iota}H^{\mathrm{BM}}_{s+t}(A_{s}^{\mathrm{trop}},A_{s-1}^{\mathrm{trop}})\xrightarrow{\partial}H_{s+t-1}^{\mathrm{BM}}(A_{s-1}^{\mathrm{trop}})\to 0,

where the middle term is isomorphic to Es,t1Q{}^{Q}\!E^{1}_{s,t}. The multiplication on E1Q{}^{Q}\!E^{1} is induced by block sum of matrices, as in Proposition 2.15, and hence coincides with (50). Therefore, for any element αHs+t1BM(As1trop;)\alpha\in H_{s+t-1}^{BM}(A_{s-1}^{\mathrm{trop}};\mathbb{Q}), Proposition 4.3 implies the following formula in E1Q{}^{Q}\!E^{1}:

(55) ι(e.ια)=ι(e).ια\iota\partial(e.\iota\alpha)=\iota\partial(e).\iota\alpha

since the second term in the Leibniz rule is e.(ιια)e.(\iota\partial\iota\alpha) which vanishes by (54). Since H1BM(A1trop)=0H_{1}^{\mathrm{BM}}(A_{1}^{\mathrm{trop}})=0, the sequence (54) reduces for (s,t)=(1,0)(s,t)=(1,0) to an isomorphism

E1,01QH0BM(A0trop;),{}^{Q}\!E^{1}_{1,0}\overset{\partial}{\cong}H_{0}^{BM}(A_{0}^{\mathrm{trop}};\mathbb{Q})\ ,

and hence by injectivity of ι\iota, we can write ι(e)=λ1E0,01Q\iota\partial(e)=\lambda 1\in{}^{Q}\!E^{1}_{0,0} for λ×\lambda\in\mathbb{Q}^{\times} and 1E0,01Q1\in{}^{Q}\!E^{1}_{0,0} the unit element in the algebra structure. In fact, by tracing through definitions, we can see λ=±1\lambda=\pm 1. The sign is not important: as in Remark 2.21, it depends on whether the chosen diffeomorphism :(0,1)(0,)\ell\colon(0,1)\to(0,\infty) in (26), appearing in the comparison of tropical and Quillen spectral sequences in subsection 2.4, is increasing or decreasing. Then (e.ια)=λα\partial(e.\iota\alpha)=\lambda\alpha, so that αλ1e.ια\alpha\mapsto\lambda^{-1}e.\iota\alpha is a right-inverse to \partial, splitting the short exact sequence (54).

Let 𝒜BM[1]\mathcal{A}^{\mathrm{BM}}[-1] denote the shift by (1,0)(-1,0) in bidegree, so 𝒜BM[1]s,t=𝒜s1,tBM.\mathcal{A}^{\mathrm{BM}}[-1]_{s,t}=\mathcal{A}^{\mathrm{BM}}_{s-1,t}. We now define the following diagram, which, we shall then argue, is commutative.

0𝒜BM𝒜BM𝒜BM[1]𝒜BM[1]00𝒜BME,1Q𝒜BM[1]0\begin{array}[]{cccccc}0\longrightarrow&\mathcal{A}^{BM}&\longrightarrow&\mathcal{A}^{BM}\oplus\mathcal{A}^{BM}[-1]&\longrightarrow&\mathcal{A}^{BM}[-1]\longrightarrow 0\\ &\downarrow&&\downarrow&&\downarrow\\ 0\longrightarrow&\mathcal{A}^{BM}&\longrightarrow&{}^{Q}\!E^{1}_{*,*}&\overset{\partial}{\longrightarrow}&\mathcal{A}^{BM}[-1]\longrightarrow 0\end{array}

Here the top row is the canonical split short exact sequence, and the bottom row is (54) in bidegree (s,t).(s,t). The vertical map in the middle is

(α,β)ια+λ1e.ιβ(\alpha,\beta)\mapsto\iota\alpha+\lambda^{-1}e.\iota\beta

where . denotes multiplication in E1Q{}^{Q}\!E^{1}, and the left and right vertical maps are the identity. By (55), (ια+λ1e.ιβ)=β\partial(\iota\alpha+\lambda^{-1}e.\iota\beta)=\beta and therefore the diagram commutes, and all vertical maps must be isomorphisms. In particular, we deduce a canonical isomorphism 𝒜BM[ϵ]/ϵ2E1Q\mathcal{A}^{BM}\otimes_{\mathbb{Q}}\mathbb{Q}[\epsilon]/\epsilon^{2}\cong{}^{Q}\!E^{1} on the level of bigraded vector spaces.

Finally, the natural morphism 𝒜BME1Q\mathcal{A}^{\mathrm{BM}}\to{}^{Q}\!E^{1} arising from the above isomorphism is a morphism of graded-commutative algebras, i.e., respects the products. Indeed, the products on domain and codomain are (49) and (50), respectively, and are both induced from the same product map (51) on perfect complexes. It follows from this, together with the definition of the middle vertical map in the above diagram, that 𝒜BM[ϵ]/ϵ2E1Q\mathcal{A}^{BM}\otimes_{\mathbb{Q}}\mathbb{Q}[\epsilon]/\epsilon^{2}\cong{}^{Q}\!E^{1} is an isomorphism of algebras. ∎

Proof of Theorem 1.1.

Any generator eE1,01Qe\in{}^{Q}\!E^{1}_{1,0}\cong\mathbb{Q} is primitive since the coproduct Δ\Delta on E1Q{}^{Q}\!E^{1} respects the bigrading (53). Therefore the ideal generated by ee is also a co-ideal, and the quotient Hopf algebra E1Q/(e){}^{Q}\!E^{1}/(e) is isomorphic to 𝒜BM\mathcal{A}^{\mathrm{BM}} by Proposition 4.4. This yields a Hopf algebra structure on 𝒜BM\mathcal{A}^{\mathrm{BM}}, and dualizing yields a Hopf algebra structure on W0Hc(𝒜)W_{0}H^{*}_{c}(\mathcal{A}). ∎

Next, we show that there is a natural injection from the bigraded vector space Ωc[1]\Omega_{c}^{*}[-1] into the subspace of primitives Prim(W0Hc(𝒜))\operatorname{Prim}(W_{0}H^{*}_{c}(\mathcal{A})). We recall the canonical differential forms for GLg\mathrm{GL}_{g} and their basic properties from §2.5.

Proof of Theorem 1.2.

By [Bro23] there is an injective map of graded \mathbb{R}-vector spaces

(56) Ωc(g)[1]Hc(Pg/GLg();)E1g,Q\Omega^{*}_{c}(g)[-1]\otimes\mathbb{R}\to H^{*}_{c}(P_{g}/\mathrm{GL}_{g}(\mathbb{Z});\mathbb{R})\cong{}^{Q}\!E_{1}^{g,\ast}\otimes\mathbb{R}

for all g>1g>1 odd (see Section 2.5 for further detail).

It remains to show that the image of Ωc(g)[1]\Omega_{c}^{*}(g)[-1] in E1Q{}^{Q}\!E_{1} consists of primitive elements. Let ωΩc(g)\omega\in\Omega^{*}_{c}(g) be homogeneous of compact type where g>1g>1 is odd. We may assume that it is of the form

ωI=ω4i1+1ω4i2+1ω4ik+1\omega^{I}=\omega^{4i_{1}+1}\wedge\omega^{4i_{2}+1}\wedge\ldots\wedge\omega^{4i_{k}+1}

where I={i1,,ik}I=\{i_{1},\ldots,i_{k}\} are distinct and 4ik+1=2g14i_{k}+1=2g-1. From (29) we deduce that

ωXYI=I=JKωXJωYK\omega^{I}_{X\oplus Y}=\sum_{I=J\cup K}\omega^{J}_{X}\otimes\omega^{K}_{Y}

where the sum is over all decompositions of II into a disjoint union of (possibly empty) sets J,KJ,K, and X,YX,Y are positive definite symmetric matrices. Suppose that XX and YY both have positive rank. Since one of ωJ\omega^{J} and ωK\omega^{K} must necessarily have ω2g1\omega^{2g-1} as a factor, it follows from (30) that ωXJωYK=0\omega^{J}_{X}\otimes\omega^{K}_{Y}=0 for all J,KJ,K and we deduce that ωXYI\omega^{I}_{X\oplus Y} is identically zero. If Δ=m,n1Δm,n\Delta^{\prime}=\sum_{m,n\geq 1}\Delta_{m,n} denotes the reduced coproduct, defined by Δ=id1+1id+Δ\Delta=\mathrm{id}\otimes 1+1\otimes\mathrm{id}+\Delta^{\prime}, then we have shown that Δω=0\Delta^{\prime}\omega=0 for all ωΩc(g)\omega\in\Omega_{c}(g) since the restriction of ω\omega to the image of 𝒜mtrop×𝒜ntrop\mathcal{A}_{m}^{\mathrm{trop}}\times\mathcal{A}_{n}^{\mathrm{trop}} for m+n=gm+n=g and m,n>0m,n>0 is zero. It follows that the (m,n)(m,n)-component of the coproduct Δω\Delta\omega is zero, and hence the image of Ωc(g)[1]\Omega_{c}(g)[-1] in E1Q{}^{Q}\!E_{1} is primitive. The theorem follows by quotienting by e.e.

Proof of Corollary 1.9.

Recall that

H(GLn();Stn)H(n2)(GLn();or)H_{*}(\mathrm{GL}_{n}(\mathbb{Z});\operatorname{St}_{n}\otimes\mathbb{Q})\cong H^{\binom{n}{2}-*}(\mathrm{GL}_{n}(\mathbb{Z});\mathbb{Q}_{\mathrm{or}})

where or\mathbb{Q}_{\mathrm{or}} denotes the GLn()\mathrm{GL}_{n}(\mathbb{Z})-module given by orientations on SLn()/SOn()\mathrm{SL}_{n}(\mathbb{R})/\mathrm{SO}_{n}(\mathbb{R}), the symmetric space of positive definite symmetric bilinear forms on n\mathbb{R}^{n} of determinant 11. To relate to SLn()\mathrm{SL}_{n}(\mathbb{Z}), Shapiro’s Lemma [Wei94, 6.3.2, p. 171] gives

H(SLn();)H(GLn();IndGLn()SLn())H^{*}(\mathrm{SL}_{n}(\mathbb{Z});\mathbb{Q})\cong H^{*}(\mathrm{GL}_{n}(\mathbb{Z});\mathrm{Ind}\begin{subarray}{c}\mathrm{GL}_{n}(\mathbb{Z})\\ \mathrm{SL}_{n}(\mathbb{Z})\end{subarray}\,\mathbb{Q})

and IndGLn()SLn()~\mathrm{Ind}\begin{subarray}{c}\mathrm{GL}_{n}(\mathbb{Z})\\ \mathrm{SL}_{n}(\mathbb{Z})\end{subarray}\,\mathbb{Q}\cong\mathbb{Q}\oplus\widetilde{\mathbb{Q}} where ~\widetilde{\mathbb{Q}} denotes the determinantal representation of GLn()\mathrm{GL}_{n}(\mathbb{Z}). Now by [EVGS13, Lemma 7.2], the orientation module or\mathbb{Q}_{\mathrm{or}} is isomorphic to \mathbb{Q} if nn is odd and to ~\widetilde{\mathbb{Q}} if nn is even. Thus, if nn is odd then GLn()SLn()×/2\mathrm{GL}_{n}(\mathbb{Z})\cong\mathrm{SL}_{n}(\mathbb{Z})\times\mathbb{Z}/2\mathbb{Z}, and

H(SLn();)H(GLn();)H(GLn();or).H^{*}(\mathrm{SL}_{n}(\mathbb{Z});\mathbb{Q})\cong H^{*}(\mathrm{GL}_{n}(\mathbb{Z});\mathbb{Q})\cong H^{*}(\mathrm{GL}_{n}(\mathbb{Z});\mathbb{Q}_{\mathrm{or}}).

If nn is even, then

H(SLn();)H(GLn();)H(GLn();~)H(GLn();)H(GLn();or).H^{*}(\mathrm{SL}_{n}(\mathbb{Z});\mathbb{Q})\cong H^{*}(\mathrm{GL}_{n}(\mathbb{Z});\mathbb{Q})\oplus H^{*}(\mathrm{GL}_{n}(\mathbb{Z});\widetilde{\mathbb{Q}})\cong H^{*}(\mathrm{GL}_{n}(\mathbb{Z});\mathbb{Q})\oplus H^{*}(\mathrm{GL}_{n}(\mathbb{Z});\mathbb{Q}_{\mathrm{or}}).

In both cases, H(GLn();or)H^{*}(\mathrm{GL}_{n}(\mathbb{Z});\mathbb{Q}_{\mathrm{or}}) is a summand of H(SLn();)H^{*}(\mathrm{SL}_{n}(\mathbb{Z});\mathbb{Q}). Now we have

(W0Hc(𝒜))[x]/x2gH(GLg(),Stg),(W_{0}H^{*}_{c}(\mathcal{A}))^{\vee}\otimes_{\mathbb{Q}}\mathbb{Q}[x]/x^{2}\cong\bigoplus_{g}H_{*}(\mathrm{GL}_{g}(\mathbb{Z}),\operatorname{St}_{g}\otimes\mathbb{Q}),

where xx has genus 11 and degree 11, from dualizing Proposition 4.4. Now Corollary 1.9 follows from Corollary 1.5. ∎

See subsection 7.3.2 for a refinement of Corollary 1.9.

4.3. Consequences of Theorem 1.2 and a conjecture

Theorem 1.2 can be rephrased as follows:

Theorem 4.5.

There is a canonical morphism of bigraded Hopf algebras

(57) T(Ωc[1])W0Hc(𝒜;),T(\Omega^{*}_{c}[-1])\otimes\mathbb{R}\longrightarrow W_{0}H^{*}_{c}(\mathcal{A};\mathbb{R}),

whose domain is given the coproduct induced by declaring Ωc[1]T(Ωc[1])\Omega^{*}_{c}[-1]\subset T(\Omega^{*}_{c}[-1]) to be primitive. The restriction of this Hopf algebra map to Ωc[1]\Omega^{*}_{c}[-1] is injective.

The map (57) restricts to a map of graded Lie algebras

(58) 𝕃(Ωc[1])Prim(W0Hc(𝒜;)),\mathbb{L}(\Omega^{*}_{c}[-1])\longrightarrow\operatorname{Prim}(W_{0}H^{*}_{c}(\mathcal{A};\mathbb{R}))\ ,

where 𝕃\mathbb{L} denotes the free (graded-commutative) Lie algebra on Ωc[1]\Omega^{*}_{c}[-1]. Conjecture 1.16 predicts that the classes from Ωc[1]\Omega^{*}_{c}[-1] satisfy no relations in W0Hc(𝒜;)W_{0}H^{*}_{c}(\mathcal{A};\mathbb{R}). We restate this as follows:

Conjecture 4.6.

The map (57), or equivalently, the map (58), is injective.

The later sections of this paper provide various kinds of evidence for this conjecture.

Remark 4.7.

A similar conjecture in [Bro21b] states that the free Lie algebra on a space isomorphic to Ωc[1]\Omega^{*}_{c}[-1] injects into the cohomology of the commutative graph complex (or equivalently, in the cohomology of the moduli spaces of tropical curves). There are some key differences between these two conjectures: the map in Conjecture 4.6 is given explicitly and respects the bigrading; the one in [Bro21b] is not explicit and does not respect the grading by genus in general. These two conjectures are not related by the tropical Torelli map: one can show that tropical Torelli map is zero on the image of almost all canonical forms which lie above the diagonal where degree equals twice the genus [Bro23, §14.6]. It is, however, expected to be an isomorphism on the Lie subalgebra generated by the ω4k+1\omega^{4k+1} which lies on the diagonal.

5. Graph spectral sequences

Our next aim will be to construct a graph spectral sequence with a Hopf algebra structure analogous to that on the Quillen spectral sequence produced in Theorem 3.18, with coproduct induced by a variant of the Connes–Kreimer coproduct on the Hopf algebra of graphs constructed in [CK98]. The spectral sequence that we produce most naturally will only be a spectral sequence of bialgebras, whose pages do not admit an antipode. For our purposes this is a very minor defect. In subsection 5.6, we will explain how to obtain a spectral sequence of connected Hopf algebras, as the localization obtained by inverting a certain element of bidegree (0,0)(0,0).

The graph spectral sequence arises from the construction of a “KK-theory of graphs,” a bisimplicial space BKGrBK_{\mathrm{Gr}} with the same structure and formal properties as BK()=|NS(Proj)|BK(\mathbb{Z})=|N_{\bullet}S_{\bullet}(\mathrm{Proj}_{\mathbb{Z}})|. The bisimplicial space we construct is rationally equivalent to a space MGrM_{\mathrm{Gr}} that may be interpreted as a space of metric graphs, which has a filtration by first Betti number. The E1E^{1}-page of the associated spectral sequence is the homology of a suitable graph complex and comes with a map of spectral sequences of bialgebras to the Quillen spectral sequence. In Section 6, we use this map to prove that the image of {ω5,ω9,,ω45}Ωc[1]\{\omega^{5},\omega^{9},\ldots,\omega^{45}\}\subset\Omega_{c}^{*}[-1]\otimes\mathbb{R} inside E1,Q{}^{Q}\!E_{1}^{*,*}\otimes\mathbb{R} generates a free Lie subalgebra of the primitives of W0Hc(𝒜;)W_{0}H^{*}_{c}(\mathcal{A};\mathbb{R}).

Remark 5.1.

J. Steinebrunner in recent talks has outlined a different construction of a spectral sequence whose E2E^{2}-page is given by the homology of the spaces Δg\Delta_{g} considered in [CGP21] and hence also related to the homology of Kontsevich’s graph complexes. His construction uses a filtration of the terminal modular \infty-operad.

Remark 5.2.

A related structure of a Hopf algebra on graphs appears in the work [KW17, GCKT20a, GCKT20b]. In particular, the graph complex as given by the bar construction [KW23] produces a co-operad structure with a multiplication which is filtered by the first Betti number. We thank Ralph Kaufmann for helpful correspondence regarding the relationship to his work with coauthors.

Remark 5.3.

These KK-theory spaces of graphs, and their corresponding chain complexes, naturally have a cubical structure reminiscent of earlier constructions of moduli spaces of metric graphs. We refer to Culler–Vogtmann’s construction of Outer Space CVn\mathrm{CV}_{n} [CV86] and its marked-point variants CVn,s\mathrm{CV}_{n,s}. These are used to study homology of the groups Out(Fn)\mathrm{Out}(F_{n}), Aut(Fn)\mathrm{Aut}(F_{n}), and a more general family of groups denoted Γn,s\Gamma_{n,s} [Hat95, HV04, CHKV16]. The space CVn\mathrm{CV}_{n} is simplicial, but it has a deformation retraction to a subcomplex 𝕊CVn\mathbb{S}\mathrm{CV}_{n} of its barycentric subdivision called the spine, which is a cubical space parametrizing marked metric graphs with a chosen subforest and has an associated graph complex [HV98, §3]. The space MGrM_{\mathrm{Gr}} that we shall consider, and its associated graph complex, have a similar structure.

Remark 5.4.

There is a Waldhausen category 𝒢\mathcal{G}, whose objects are all connected graphs GG together with chosen basepoint V(G)\ast\in V(G), whose morphisms are all maps of graphs preserving base point, cofibrations the injective maps, and weak equivalences the ones whose corresponding map of topological spaces is a homotopy equivalence. There is also a symmetric monoidal structure given by wedge sum. Filtering by b1(G)b_{1}(G)\in\mathbb{N} induces a filtration of BK(𝒢)BK(\mathcal{G}) with the exact same formal properties as the filtration on BK()BK(\mathbb{Z}), leading to a spectral sequence of Hopf algebras

Es,t1Hs+t(BK(𝒢)).E^{1}_{s,t}\Rightarrow H_{s+t}(BK(\mathcal{G})).

Moreover, the functor H1:𝒢ProjH_{1}\colon\mathcal{G}\to\mathrm{Proj}_{\mathbb{Z}} taking GG to H1(G;)H_{1}(G;\mathbb{Z}) preserves all the structure and induces a map of spectral sequences of Hopf algebras.

At present we do not know how to make good use of this specific spectral sequence. Instead we proceed below with a variant involving disconnected graphs without basepoint. It does not seem to literally fit into the axioms of a Waldhausen category, but we will explain an explicit construction which seems similar in spirit, and which we shall therefore denote BKGrBK_{\mathrm{Gr}}.

5.1. KK-theory of graphs

We use the same definition as in [CGP21]: a graph is a finite set XX, together with functions i,r:XXi,r\colon X\to X that satisfy i2=1Xi^{2}=1_{X} and r2=rr^{2}=r, and such that

{xXr(x)=x}={xXi(x)=x}.\{x\in X\mid r(x)=x\}=\{x\in X\mid i(x)=x\}.

We write V={xXi(x)=x}V=\{x\in X\mid i(x)=x\}, H=XVH=X\setminus V, and E=H/(xi(x))E=H/(x\sim i(x)), for the sets of vertices, half-edges, and edges of GG, respectively. The most general notion of morphism from G=(X,i,r)G=(X,i,r) to G=(X,i,r)G^{\prime}=(X^{\prime},i^{\prime},r^{\prime}) is just a map of sets ϕ:XX\phi\colon X\to X^{\prime} such that iϕ=ii^{\prime}\circ\phi=i and rϕ=rr^{\prime}\circ\phi=r; we do not require that it send HH to HH^{\prime}. When we impose extra conditions on graphs (non-empty, connected, no zero-valent vertices, etc.) or graph morphisms, we say so explicitly. A morphism of graphs is injective (resp. surjective) if the underlying map on sets is injective (resp. surjective).

While officially there are no extra conditions on morphisms, in practice only a restricted class of morphisms will appear in the diagrams of graphs below. For example, surjective morphisms of graphs can in principle increase first Betti number by identifying vertices (for example, one may map the two endpoints of an edge to a single vertex, to create a self-edge). But such a morphism will never arise in the diagrams of graphs considered below.

Since we allow vertices of any valence including 0, the category of finite sets may be identified via S(S,id,id)S\mapsto(S,\mathrm{id},\mathrm{id}) with a full subcategory of the category of graphs, namely those graphs GG for which E(G)=E(G)=\emptyset. Any graph GG comes with universal maps to and from a set (more precisely, a map of graphs from GG to a graph with no edges, initial among such, as well as a map of graphs from a graph with no edges to GG, terminal among such) namely

V(G)Gπ0(G),V(G)\hookrightarrow G\twoheadrightarrow\pi_{0}(G),

where V(G)V(G) is the set of vertices and π0(G)\pi_{0}(G) is the set of path components of GG.

We now construct a KK-theory space of graphs, corresponding to a simplicial category \mathcal{F} that we now define. For an object [p]Δ[p]\in\Delta, p\mathcal{F}_{p} is a groupoid whose object set is the set of diagrams of the shape

(59) G0,0\textstyle{G_{0,0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}G0,1\textstyle{G_{0,1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}G0,2\textstyle{G_{0,2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}G0,p\textstyle{G_{0,p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}G1,1\textstyle{G_{1,1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}G1,2\textstyle{G_{1,2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}G1,p\textstyle{G_{1,p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\vdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\vdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Gp1,p1\textstyle{G_{p-1,p-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Gp1,p\textstyle{G_{p-1,p}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Gp,p\textstyle{G_{p,p}}

of graphs Gi,jG_{i,j} and morphisms between them, subject to the conditions

  • E(Gi,i)=E(G_{i,i})=\emptyset for i=0,,pi=0,\dots,p,

  • the maps V(Gi,j)V(Gi,j+1)V(G_{i,j})\to V(G_{i,j+1}) induced by the horizontal maps in the diagram are bijections, and E(Gi,j)E(Gi,j+1)E(G_{i,j})\to E(G_{i,j+1}) are injections, 0ij<p0\leq i\leq j<p,

  • the maps π0(Gi1,j)π0(Gi,j)\pi_{0}(G_{i-1,j})\to\pi_{0}(G_{i,j}) induced by the vertical maps are bijections,

  • all squares in the diagrams are pushout diagrams.

The arrows denoted \rightarrowtail in the diagram (59) are the ones that are required to be bijections on vertices and injections on edges. The conditions ensure that every vertical map is induced by contraction of edges only. The morphisms in p\mathcal{F}_{p} are isomorphisms of such diagrams. Deleting the iith row and column define functors di:pp1d_{i}\colon\mathcal{F}_{p}\to\mathcal{F}_{p-1} for 0ip0\leq i\leq p, and there are also functors si:pp+1s_{i}\colon\mathcal{F}_{p}\to\mathcal{F}_{p+1} defined by inserting identities, making [p]p[p]\mapsto\mathcal{F}_{p} a simplicial object in the category of small groupoids (see Remark 5.6).

Remark 5.5.

  1. (1)

    The diagonal maps Gi1,i1Gi,iG_{i-1,i-1}\to G_{i,i} are surjections of finite sets. The diagram provides a factorization

    Gi1,i1=V(Gi1,i1)V(Gi1,i)π0(Gi1,i)π0(Gi,i)Gi,iG_{i-1,i-1}=V(G_{i-1,i-1})\cong V(G_{i-1,i})\twoheadrightarrow\pi_{0}(G_{i-1,i})\cong\pi_{0}(G_{i,i})\cong G_{i,i}

    identifying Gi1,i1Gi,iG_{i-1,i-1}\twoheadrightarrow G_{i,i} with the quotient by the equivalence relation generated by the edge set of Gi1,iG_{i-1,i}.

  2. (2)

    The diagram (59) is determined up to isomorphism by its top row; since each rectangle

    Gi1,i\textstyle{G_{i-1,i}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Gi1,j\textstyle{G_{i-1,j}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Gi,i\textstyle{G_{i,i}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Gi,j\textstyle{G_{i,j}}

    is a pushout diagram, each vertical map Gi1,jGi,jG_{i-1,j}\to G_{i,j} must be induced by collapsing each edge in the image of Gi1,iGi1,jG_{i-1,i}\hookrightarrow G_{i-1,j}. The whole diagram can therefore be reconstructed up to isomorphism from the graph G0,pG_{0,p} together with the flag of subsets of E(G0,p)E(G_{0,p}) given by the images of E(G0,j)E(G0,p)E(G_{0,j})\hookrightarrow E(G_{0,p}). As in the Waldhausen construction, it is, for set-theoretic reasons, convenient to include the data of chosen subquotients.

  3. (3)

    Similarly, the diagram is determined up to isomorphism by its rightmost column.

We define

(60) BKGr=|N|.BK_{\mathrm{Gr}}=|N_{\bullet}\mathcal{F}_{\bullet}|.

There is a functor

pH1Sp(Proj),Gi,jH1(Gi,j;),\mathcal{F}_{p}\stackrel{{\scriptstyle H_{1}}}{{\longrightarrow}}S_{p}(\mathrm{Proj}_{\mathbb{Z}}),\qquad G_{i,j}\mapsto H_{1}(G_{i,j};\mathbb{Z}),

which induces a map of bisimplicial sets and in turn of topological spaces

(61) N\displaystyle N_{\bullet}\mathcal{F}_{\bullet} H1NS(Proj),\displaystyle\stackrel{{\scriptstyle H_{1}}}{{\longrightarrow}}N_{\bullet}S_{\bullet}(\mathrm{Proj}_{\mathbb{Z}}),
(62) BKGr=|N|\displaystyle BK_{\mathrm{Gr}}=|N_{\bullet}\mathcal{F}_{\bullet}| |NS(Proj)|=BK().\displaystyle\longrightarrow|N_{\bullet}S_{\bullet}(\mathrm{Proj}_{\mathbb{Z}})|=BK(\mathbb{Z}).
Remark 5.6.

For set-theoretic reasons, the above definitions should be augmented with a choice of small category 𝒢\mathcal{G} equivalent to all graphs and all morphisms (for instance by insisting G=(X,i,r)G=(X,i,r) where XΩX\subset\Omega is a subset of some fixed infinite set, or by choosing a graph of each isomorphism type). Then we get a small category p\mathcal{F}_{p} whose object set is the set of diagrams in 𝒢\mathcal{G} of the form (59), subject to the stated requirements.

5.2. Filtrations and graph spectral sequences

Definition 5.7.

For gg\in\mathbb{N}, let FgppF_{g}\mathcal{F}_{p}\subset\mathcal{F}_{p} be the full subcategory containing those objects (59) with b1(G0,p)gb_{1}(G_{0,p})\leq g. Similarly, let FgBKGrBKGrF_{g}BK_{\mathrm{Gr}}\subset BK_{\mathrm{Gr}} be the image of the map induced by N(Fg)NN_{\bullet}(F_{g}\mathcal{F}_{\bullet})\to N_{\bullet}\mathcal{F}_{\bullet}, the nerve of the inclusion of the subcategory.

We have now defined filtrations on both spaces in (62), and it is clear that the map preserves the filtration. Therefore there is an induced map of spectral sequences

(63) E,rGE,rQ,{}^{G}\!E^{r}_{*,*}\to{}^{Q}\!E^{r}_{*,*},

converging to the rational homology of the spaces (62). We will call E,rG{}^{G}\!E^{r}_{*,*} the “graph spectral sequence.”

Proposition 5.8.

The constructions in Section 3 apply to the bisimplicial set Xp,q=NqpX_{p,q}=N_{q}\mathcal{F}_{p} filtered as FgXp,q=Nq(Fgp)F_{g}X_{p,q}=N_{q}(F_{g}\mathcal{F}_{p}), leading to a space-level filtered coproduct

|X||es(X)||X|×|X|.|X|\xrightarrow{\approx}|\mathrm{es}(X)|\to|X|\times|X|.

Choosing a disjoint union operation (G,G)m(G,G)GG(G,G^{\prime})\mapsto m(G,G^{\prime})\cong G\sqcup G^{\prime} on the chosen small category equivalent to all finite graphs and promoting it to a symmetric monoidal functor leads to a map of bisimplicial sets X×XXX\times X\to X inducing a product

m:|X|×|X||X×X||X|.m\colon|X|\times|X|\xrightarrow{\approx}|X\times X|\to|X|.

The product and coproduct are both filtered maps, the product is associative and commutative up to a filtration-preserving homotopy, and the coproduct is associative up to a filtration-preserving homotopy. Forgetting filtrations, the coproduct is homotopic to the diagonal map of |X||X|.

Corollary 5.9.

Let E,1GH(BKGr){}^{G}\!E^{1}_{*,*}\Rightarrow H_{*}(BK_{\mathrm{Gr}}) be the homological spectral sequence associated to the filtration on BKGrBK_{\mathrm{Gr}} defined above. Then E,G{}^{G}\!E^{*}_{*,*} admits the structure of a spectral sequence of bialgebras, with graded commutative product on all pages and graded co-commutative co-product on EE^{\infty}.

As we will see later, the graph spectral sequence will have E0,01[x]/(x2x)E^{1}_{0,0}\cong\mathbb{Q}[x]/(x^{2}-x), with unit 11 the class of the empty graph and xE0,01x\in E^{1}_{0,0} the class of a graph consisting of a single zero-valent vertex. In this bidegree the coproduct is given by Δ(x)=xx\Delta(x)=x\otimes x and the augmentation by x1x\mapsto 1. This bigraded bialgebra is not connected and the existence of an antipode is not automatic, hence “bialgebra” instead of “Hopf algebra” in the above statement. (And indeed it does not admit an antipode: recall that Spec\mathrm{Spec} of a commutative bialgebra is a monoid scheme, and is a group scheme if and only if the bialgebra is part of a Hopf algebra structure. See, e.g., [Mil12, Theorem 5.1]. The bialgebra E0,01E^{1}_{0,0} represents the functor sending a commutative ring RR to its monoid of idempotent elements—this defines a monoid scheme Spec([x]/(x2x))\operatorname{Spec}(\mathbb{Q}[x]/(x^{2}-x)) which is not a group scheme.)

Proof sketch for Proposition 5.8.

The coproduct is constructed just as for BK()BK(\mathbb{Z}), using the filtration and the bisimplicial structure. The coassociativity of the coproduct, up to a homotopy respecting the filtration, follows from Proposition 3.11 and the observation that given

G0G2p+1,G_{0}\rightarrowtail\cdots\rightarrowtail G_{2p+1},

where the arrows \rightarrowtail are bijections on vertices and injections on edges, the inequality

b1(G2p+1)b1(Gp)+b1(G2p+1/Gp+1)b_{1}(G_{2p+1})\geq b_{1}(G_{p})+b_{1}(G_{2p+1}/G_{p+1})

holds. From this observation, it follows that

es(Fs(NF))Fs(NF×NF)Fs(NF)×Fs(NF),\mathrm{es}(F_{s}(N_{\bullet}F_{\bullet}))\Rightarrow F_{s}(N_{\bullet}F_{\bullet}\times N_{\bullet}F_{\bullet})\Rightarrow F_{s}(N_{\bullet}F_{\bullet})\times F_{s}(N_{\bullet}F_{\bullet}),

so that the hypotheses of Proposition 3.11 are satisfied.

The product on Sp(Proj)S_{p}(\mathrm{Proj}_{\mathbb{Z}}) came from the symmetric monoidal structure induced by direct sum, but we can define a symmetric monoidal structure on p\mathcal{F}_{p} by disjoint union of graphs, and clearly H1:pSp(Proj)H_{1}\colon\mathcal{F}_{p}\to S_{p}(\mathrm{Proj}_{\mathbb{Z}}) promotes to a symmetric monoidal functor. ∎

In the remainder of this section, we explain why a filtered space rationally equivalent to BKGrBK_{\mathrm{Gr}} may be interpreted as a moduli space of possibly-disconnected metric graphs, and explain how the E1E^{1}-page may be identified with homology of a certain graph complex.

Definition 5.10.

Let ¯p\overline{\mathcal{F}}_{p} denote the set of isomorphism classes in the groupoid p\mathcal{F}_{p}, in other words the coequalizer of d0,d1:N1pN0pd_{0},d_{1}\colon N_{1}\mathcal{F}_{p}\to N_{0}\mathcal{F}_{p} in the category of sets, or the set π0(|Np|)\pi_{0}(|N_{\bullet}\mathcal{F}_{p}|) of path components of the geometric realization of the nerve of p\mathcal{F}_{p}.

The sets ¯p\overline{\mathcal{F}}_{p} assemble into a simplicial set ¯\overline{\mathcal{F}}_{\bullet}. We set the notation

(64) MGr=|¯|,M_{\mathrm{Gr}}=|\overline{\mathcal{F}}_{\bullet}|,

since, as explained in Section 5.3 below, this space has an interpretation as a coarse moduli space of graphs. Regarding ¯p\overline{\mathcal{F}}_{p} as a space with discrete topology, the canonical maps

(65) |Np|¯p|N_{\bullet}\mathcal{F}_{p}|\to\overline{\mathcal{F}}_{p}

for all pp assemble to a map of simplicial spaces, with geometric realization

(66) BKGr=|N||¯|=MGr.BK_{\mathrm{Gr}}=|N_{\bullet}\mathcal{F}_{\bullet}|\to|\overline{\mathcal{F}}_{\bullet}|=M_{\mathrm{Gr}}.

The space MGrM_{\mathrm{Gr}} does not map to BK()BK(\mathbb{Z}), but is a simpler object than BKGrBK_{\mathrm{Gr}} in that it arises from a simplicial rather than bisimplicial set. Furthermore, it is a good model up to rational equivalence:

Lemma 5.11.

The maps (65) and (66) induce isomorphisms in rational homology. Filtering ¯p\overline{\mathcal{F}}_{p} by the images of the filtration in N0pN_{0}\mathcal{F}_{p}, the map (66) also induces an isomorphism in rational homology of associated gradeds.

Proof.

The homotopy type of |Np||N_{\bullet}\mathcal{F}_{p}| is a disjoint union of K(π,1)K(\pi,1)-spaces for the automorphism groups of diagrams of the form (59), one for each isomorphism class of such diagrams. Since any finite group has the rational homology of a point, the map (65) is a rational equivalence for each pp. Filtering the realization in the pp-direction by skeletons shows that (66) is too. ∎

5.3. Rational cell decomposition of graph spaces

The geometric realization MGrM_{\mathrm{Gr}} can be described rather explicitly, as an iterated “rational cell attachment” in which each cell is the cone over the quotient of a sphere by a finite group, i.e., a symmetric CW-complex as defined in [ACP22]. The cells here most naturally have cubical shape, attaching the cube (Δ1)n(\Delta^{1})^{n} to a space XX along an attaching map e:(Δ1)nXe\colon\partial(\Delta^{1})^{n}\to X, and more generally as a pushout of the form

(Δ1)n/H(Δ1)n/H𝑒X(\Delta^{1})^{n}/H\leftarrow\partial(\Delta^{1})^{n}/H\xrightarrow{e}X

for a subgroup H<SnH<S_{n} acting by permuting coordinates. When HAnH\leq A_{n} then (Δ1)n/H\partial(\Delta^{1})^{n}/H is a rational homology Sn1S^{n-1} and the attachment has the same effect in rational homology as attaching an ordinary nn-cell in the usual sense. When HH contains an odd permutation, then the attachment does not change the rational homology. As mentioned in Remark 5.3, this description of MGrM_{\mathrm{Gr}} as a cubical space of graphs, and the graph complex we shall associate to it in Section 5.4, is a variant of the cubical spaces of graphs introduced by Hatcher–Vogtmann [HV98, §3] in their work computing rational homology groups of Aut(Fn)\operatorname{Aut}(F_{n}). It is also similar in spirit, although not isomorphic, to the notion of “symmetric Δ\Delta-complex” in [CGP21]: in that paper an iterated attachment of the pair Δn1/HΔn1/H\partial\Delta^{n-1}/H\subset\Delta^{n-1}/H was used, a rational homology (n1)(n-1)-disk, instead of the pair (Δ1)n/H(Δ1)n/H\partial(\Delta^{1})^{n}/H\subset(\Delta^{1})^{n}/H, a rational homology nn-disk.

Observe that any graph GG gives rise to a map of simplicial sets

(Δ1)E(G)N0()¯,(\Delta^{1}_{\bullet})^{E(G)}\to N_{0}(\mathcal{F}_{\bullet})\to\overline{\mathcal{F}}_{\bullet},

where Δ1\Delta^{1}_{\bullet} is the usual representable simplicial set [p]Δ([p],[1])[p]\mapsto\Delta([p],[1]) with geometric realization Δ1\Delta^{1}. We now explain this map precisely. Given a pp-simplex (fe)eE(G)(f_{e})_{e\in E(G)} of (Δ1)E(G)(\Delta^{1}_{\bullet})^{E(G)}, where feΔ([p],[1])f_{e}\in\Delta([p],[1]), we associate the sequence of subsets of E(G)E(G)

(67) E0EpE(G)\emptyset\subset E_{0}\subset\dots\subset E_{p}\subset E(G)

in which

(68) Ej={eE(G)fe(j)=1}.E_{j}=\{e\in E(G)\mid f_{e}(j)=1\}.

We may now form a diagram of graphs of the form (59), by first writing G1,jG_{-1,j} for the graph with V(G1,j)=V(G)V(G_{-1,j})=V(G) and E(G1,j)=EjE(G_{-1,j})=E_{j}, and then for 0ijp0\leq i\leq j\leq p defining Gi,jG_{i,j} as the quotient of G1,jG_{-1,j} obtained by collapsing those edges eE(G1,j)=Eje\in E(G_{-1,j})=E_{j} which are elements of the subset EiEjE_{i}\subset E_{j}. This recipe gives a diagram of the form (59) satisfying all requirements, except that the chosen quotient graphs Gi,jG_{i,j} may not literally be in the object set of 𝒢\mathcal{G}. (Recall that we chose a small category in order for NN_{\bullet}\mathcal{F}_{\bullet} to be a bisimplicial set.) Each such diagram is certainly isomorphic to a diagram in 𝒢\mathcal{G}, and choosing an isomorphic diagram in 𝒢\mathcal{G} for each pp and each fΔp1f\in\Delta^{1}_{p} produces a map of simplicial sets (Δ1)E(G)N0(\Delta^{1}_{\bullet})^{E(G)}\to N_{0}\mathcal{F}_{\bullet}. The map (Δ1)E(G)N0()(\Delta^{1}_{\bullet})^{E(G)}\to N_{0}(\mathcal{F}_{\bullet}) constructed in this way depends on choices, but the composition (Δ1)E(G)N0()¯(\Delta^{1}_{\bullet})^{E(G)}\to N_{0}(\mathcal{F}_{\bullet})\to\overline{\mathcal{F}}_{\bullet} does not, and factors over the quotient (Δ1)E(G)/Aut(G)(\Delta^{1}_{\bullet})^{E(G)}/\mathrm{Aut}(G). This gives commutative diagrams

(Δ1)E(G){(\Delta^{1}_{\bullet})^{E(G)}}N0(){N_{0}(\mathcal{F}_{\bullet})}(Δ1)E(G)/Aut(G){(\Delta^{1}_{\bullet})^{E(G)}/\mathrm{Aut}(G)}¯.{\overline{\mathcal{F}}_{\bullet}.}

Recall that the geometric realization functor preserves small colimits and finite products, and that |Δ1|=Δ1|\Delta^{1}_{\bullet}|=\Delta^{1}, so the induced maps of geometric realizations may be written as

(69) (Δ1)E(G)/Aut(G)|(Δ1)E(G)/Aut(G)|MGr.(\Delta^{1})^{E(G)}/\mathrm{Aut}(G)\xleftarrow{\approx}|(\Delta^{1}_{\bullet})^{E(G)}/\mathrm{Aut}(G)|\longrightarrow M_{\mathrm{Gr}}.

The following lemma summarizes the sense in which these canonical maps (Δ1)E(G)/Aut(G)MGr(\Delta^{1})^{E(G)}/\mathrm{Aut}(G)\to M_{\mathrm{Gr}} behave (rationally) like cells of a CW structure. It also defines an analogue of the filtration by skeletons (not to be confused with the more interesting filtration by first Betti number, which leads to the graph spectral sequence).

Lemma 5.12.

Let ¯(n)¯\overline{\mathcal{F}}^{(n)}_{\bullet}\subset\overline{\mathcal{F}}_{\bullet} denote the simplicial subset whose pp-simplices are cut out by the condition that the cardinality of E(G0,p)E(G_{0,p}) is at most nn. Then the maps (69) assemble to pushout diagrams of topological spaces

G(Δ1)E(G)/Aut(G){\coprod_{G}\partial(\Delta^{1})^{E(G)}/\mathrm{Aut}(G)}|¯(n1)|{{\big{|}\overline{\mathcal{F}}_{\bullet}^{(n-1)}\big{|}}}G(Δ1)E(G)/Aut(G){\coprod_{G}(\Delta^{1})^{E(G)}/\mathrm{Aut}(G)}|¯(n)|,{{\big{|}\overline{\mathcal{F}}_{\bullet}^{(n)}\big{|}},}

where the coproduct is indexed by graphs GG with |E(G)|=n|E(G)|=n, one in each isomorphism class.

The entry (Δ1)E(G)/Aut(G)\partial(\Delta^{1})^{E(G)}/\mathrm{Aut}(G) in the diagram should be parsed as (((Δ1)E(G)))/Aut(G)\big{(}\partial\big{(}(\Delta^{1})^{E(G)}\big{)}\big{)}/\mathrm{Aut}(G): the orbit space of the action of Aut(G)\mathrm{Aut}(G) on the boundary of the cube (Δ1)E(G)(\Delta^{1})^{E(G)}.

Using this presentation as an iterated pushout, we deduce descriptions of the associated graded with respect to this filtration of MGrM_{\mathrm{Gr}} by number of edges, as well as a description of the underlying point set.

Corollary 5.13.

  1. (1)

    The maps (69) induce homeomorphisms of pointed spaces

    |E(G)|=n(S1)E(G)/Aut(G)|¯(n)||¯(n1)|\bigvee_{|E(G)|=n}(S^{1})^{\wedge E(G)}/\mathrm{Aut}(G)\longrightarrow\frac{|\overline{\mathcal{F}}^{(n)}_{\bullet}|}{|\overline{\mathcal{F}}^{(n-1)}_{\bullet}|}

    where the coproduct of pointed spaces is indexed by graphs with precisely nn edges, one in each isomorphism class, and Sn(S1)E(G)S^{n}\approx(S^{1})^{\wedge E(G)} is the E(G)E(G)-fold smash product of the circle with itself, on which Aut(G)\mathrm{Aut}(G) acts by permuting factors.

  2. (2)

    The maps (69) induce a continuous bijection

    G(Δ1Δ1)E(G)/Aut(G)MGr,\coprod_{G}(\Delta^{1}\setminus\partial\Delta^{1})^{E(G)}/\mathrm{Aut}(G)\longrightarrow M_{\mathrm{Gr}},

    where the disjoint union is over all graphs, one in each isomorphism class.

Remark 5.14.

The second part of the corollary shows that the set underlying MGrM_{\mathrm{Gr}} may be interpreted as the set of isomorphism classes of pairs (G,)(G,\ell) where :E(G)Δ1Δ1\ell\colon E(G)\to\Delta^{1}\setminus\partial\Delta^{1} is any function, and [G,]=[G,][G,\ell]=[G^{\prime},\ell^{\prime}] if there exists an isomorphism ϕ:GG\phi\colon G\to G^{\prime} such that (ϕ(e))=(e)\ell^{\prime}(\phi(e))=\ell(e) for all eE(G)e\in E(G). For psychological reasons it may be preferable to rescale using a homeomorphism Δ1Δ1(0,)\Delta^{1}\setminus\partial\Delta^{1}\to(0,\infty) such as (t0,t1)log(t1)(t_{0},t_{1})\mapsto-\log(t_{1}). This gives an interpretation of MGrM_{\mathrm{Gr}} as a coarse moduli space of metric graphs, where edges have assigned positive real lengths. In this picture, the topology may described informally as “edges are collapsed as their lengths go to zero, and edges are deleted as their lengths go to infinity.”

Proof of Lemma 5.12.

Inspecting the definition of the canonical map (69), we see that it factors through the subspace |¯(n)||\overline{\mathcal{F}}^{(n)}_{\bullet}| when |E(G)|n|E(G)|\leq n and that for each eE(G)e\in E(G), its composition with

(Δ1){e}×(Δ1)E(G){e}(Δ1)E(G)(Δ1)E(G)/Aut(G)(\partial\Delta^{1})^{\{e\}}\times(\Delta^{1})^{E(G)\setminus\{e\}}\hookrightarrow(\Delta^{1})^{E(G)}\twoheadrightarrow(\Delta^{1})^{E(G)}/\mathrm{Aut}(G)

will factor through |¯(n1)||\overline{\mathcal{F}}^{(n-1)}_{\bullet}| because the only graphs that will appear are subquotients of GeG\setminus e and G/eG/e, so they will have strictly fewer edges than GG has. This observation explains why (69) induce horizontal maps in the square diagram in the lemma, and it is then clear that the diagram commutes.

To prove that the diagram is a pushout, it suffices to consider the corresponding diagram of simplicial sets, again because geometric realization preserves small colimits. Pushouts in simplicial sets are computed degree-wise, so it suffices to prove that the diagram of pp-simplices

G((Δ1)E(G))p/Aut(G){\coprod_{G}(\partial(\Delta^{1}_{\bullet})^{E(G)})_{p}/\mathrm{Aut}(G)}¯p(n1){{\overline{\mathcal{F}}_{p}^{(n-1)}}}G(Δp1)E(G)/Aut(G){\coprod_{G}(\Delta^{1}_{p})^{E(G)}/\mathrm{Aut}(G)}¯p(n){{\overline{\mathcal{F}}_{p}^{(n)}}}

is a pushout in sets. Both vertical maps are injective. We must show that the horizontal map restricts to a bijection between the complements of the images of the vertical maps.

Before passing to Aut(G)\mathrm{Aut}(G)-orbits, the subset ((Δ1)E(G))p(Δp1)E(G)=Δ([p],[1])E(G)(\partial(\Delta^{1}_{\bullet})^{E(G)})_{p}\subset(\Delta^{1}_{p})^{E(G)}=\Delta([p],[1])^{E(G)} corresponds to the set of poset maps [p][1]E(G)[p]\to[1]^{E(G)} where at least one coordinate is constant. The complement is therefore precisely the set

(Δp1)E(G)((Δ1)E(G))p=Δsurj([p],[1])E(G)(\Delta^{1}_{p})^{E(G)}\setminus(\partial(\Delta^{1}_{\bullet})^{E(G)})_{p}=\Delta_{\mathrm{surj}}([p],[1])^{E(G)}

consisting of E(G)E(G)-tuples of morphisms [p][1][p]\to[1] in the simplex category Δ\Delta whose underlying map of sets is surjective.

The proof is concluded by firstly observing that the set map (Δp1)E(G)¯p(\Delta^{1}_{p})^{E(G)}\to\overline{\mathcal{F}}_{p} defined above will send Δsurj([p],[1])E(G)\Delta_{\mathrm{surj}}([p],[1])^{E(G)} into ¯p(n)¯p(n1)\overline{\mathcal{F}}^{(n)}_{p}\setminus\overline{\mathcal{F}}^{(n-1)}_{p} if |E(G)|=n|E(G)|=n: indeed, in this case the flag of subsets (67) of E(G)E(G) will have E0=E_{0}=\emptyset and Ep=E(G)E_{p}=E(G) and hence G0,p=GG_{0,p}=G. Secondly, the resulting map of sets

|E(G)|=nΔsurj([p],[1])E(G)/Aut(G)¯p(n)¯p(n1)\coprod_{|E(G)|=n}\Delta_{\mathrm{surj}}([p],[1])^{E(G)}/\mathrm{Aut}(G)\to\overline{\mathcal{F}}^{(n)}_{p}\setminus\overline{\mathcal{F}}^{(n-1)}_{p}

is surjective because any diagram of the form (59) with |E(G0,p)|=n|E(G_{0,p})|=n arises from G=G0,pG=G_{0,p} and the flag of subsets =E0Ep=E(G)\emptyset=E_{0}\subset\dots\subset E_{p}=E(G) such that E(G0,j)=EjE(G_{0,j})=E_{j}, corresponding to some unique f=(fe)eE(G)Δsurj([p],[1])E(G)f=(f_{e})_{e\in E(G)}\in\Delta_{\mathrm{surj}}([p],[1])^{E(G)}. Conversely the map is injective because this flag of subsets of E(G)E(G) is uniquely determined up to automorphisms of GG. ∎

5.4. A graph complex modeling C(MGr)C_{*}(M_{\mathrm{Gr}})

If GG is a graph with |E(G)|=n|E(G)|=n then any choice of bijection ω:{0,,n1}E(G)\omega\colon\{0,\dots,n-1\}\to E(G) induces a homeomorphism Sn=(S1)n(S1)E(G)S^{n}=(S^{1})^{\wedge n}\approx(S^{1})^{\wedge E(G)} and induces a group homomorphism Aut(G)Sn\mathrm{Aut}(G)\to S_{n}. Let us write H=HG,ω<SnH=H_{G,\omega}<S_{n} for the image of this group homomorphism. Then ω\omega also induces a homeomorphism

(S1)E(G)/Aut(G)Sn/H(S^{1})^{\wedge E(G)}/\mathrm{Aut}(G)\xrightarrow{\cong}S^{n}/H

and the quotient map SnSn/HS^{n}\to S^{n}/H is a rational homology isomorphism if H<AnH<A_{n}, whereas Sn/HS^{n}/H has the rational homology of a point when HH contains an odd permutation. In the eyes of rational homology, the filtration of MGrM_{\mathrm{Gr}} by number of edges therefore behaves like a CW structure whose nn-cells are in bijection with the set of isomorphism classes of graph GG with |E(G)|=n|E(G)|=n and such that GG does not admit any automorphisms inducing an odd permutation of E(G)E(G). This leads to a “cellular chain complex” quasi-isomorphic to Csing(MGr;)C_{*}^{\mathrm{sing}}(M_{\mathrm{Gr}};\mathbb{Q}). This cellular chain complex has a combinatorial description which we spell out, as well as the filtration corresponding to first Betti number. The latter gives a “graph complex” point of view on the spectral sequence of bialgebras mapping to the Quillen spectral sequence.

Let C=(C,)C=(C_{*},\partial) be the rational chain complex generated by symbols [G,ω][G,\omega], where GG is any graph and ω\omega is a total order on E(G)E(G), subject to the usual relation [G,ω]=sgn(σ)[G,ω][G,\omega]=\mathrm{sgn}(\sigma)[G^{\prime},\omega^{\prime}] for each isomorphism GGG\to G^{\prime} with respect to which ω\omega and ω\omega^{\prime} are related by permutation σ\sigma. We do not impose any conditions whatsoever on G=(X,i,r)G=(X,i,r): in particular, we allow disconnected graphs, bridges, loops, parallel edges, and vertices of any valence including zero.

The degree of the generator [G,ω][G,\omega] is declared to be |E(G)||E(G)|, and we write CpCC_{p}\subset C for the span of those [G,ω][G,\omega] for which |E(G)|=p|E(G)|=p. We will return to the boundary map shortly, but first we define the comparison map to Csing(MGr;).C_{*}^{\mathrm{sing}}(M_{\mathrm{Gr}};\mathbb{Q}). To this end, recall that the identity map of Δ1\Delta^{1}, regarded as a chain ι1C1(Δ1;)\iota_{1}\in C_{1}(\Delta^{1};\mathbb{Z}), represents the fundamental class in H1(Δ1,Δ1;)H_{1}(\Delta^{1},\partial\Delta^{1};\mathbb{Z}). Then we use the cross product ([Hat02, p.277–278]) natural transformation Cp(X)Cq(Y)Cp+q(X×Y)C_{p}(X)\otimes C_{q}(Y)\to C_{p+q}(X\times Y) to define a system of fundamental chains

(70) ιnCn((Δ1)n;),\iota_{n}\in C_{n}((\Delta^{1})^{n};\mathbb{Z}),

compatible in the sense that ιn×ιm=ιn+m\iota_{n}\times\iota_{m}=\iota_{n+m}. There is also an explicit formula for ιn\iota_{n} as a signed sum of n!n! many maps Δn(Δ1)n\Delta^{n}\to(\Delta^{1})^{n}, corresponding to all the non-degenerate nn-simplices of (Δ1)n(\Delta^{1}_{\bullet})^{n} with appropriate signs, but we will not need this explicit formula.

Definition 5.15.

For a graph GG with |E(G)|=n|E(G)|=n and a total ordering ω:{0,,n1}E(G)\omega\colon\{0,\dots,n-1\}\to E(G), write fG:(Δ1)EMGrf^{G}\colon(\Delta^{1})^{E}\to M_{\mathrm{Gr}} for the canonical map defined as in (69), and fG,ω:(Δ1)nMGrf^{G,\omega}\colon(\Delta^{1})^{n}\to M_{\mathrm{Gr}} for the composition

(Δ1)n(Δ1)E(G)fGMGr(\Delta^{1})^{n}\approx(\Delta^{1})^{E(G)}\xrightarrow{f^{G}}M_{\mathrm{Gr}}

with the homeomorphism induced by the bijection ω\omega. Then define the linear map

(71) Cp\displaystyle C_{p} Cpsing(MGr;),\displaystyle\to C_{p}^{\mathrm{sing}}(M_{\mathrm{Gr}};\mathbb{Q}),
[G,ω]\displaystyle[G,\omega] fG,ω(ιn).\displaystyle\mapsto f^{G,\omega}_{*}(\iota_{n}).

Filtering by the number of edges, we deduce the following from Corollary 5.13.

Lemma 5.16.

The relative homology

H(|¯(p)|,|¯(p1)|;)H_{*}\big{(}\big{|}\overline{\mathcal{F}}^{(p)}_{\bullet}\big{|},\big{|}\overline{\mathcal{F}}^{(p-1)}_{\bullet}\big{|};\mathbb{Q}\big{)}

vanishes for p*\neq p. If |E(G)|=p|E(G)|=p, then the homomorphism (71) sends [G,ω][G,\omega] into the subspace Cp(|¯(p)|;)C_{p}(|\overline{\mathcal{F}}^{(p)}_{\bullet}|;\mathbb{Q}). Its image in relative chains

fG,ω(ιn)Cp(|¯(p)|,|¯(p1)|;)f^{G,\omega}_{*}(\iota_{n})\in C_{p}\big{(}\big{|}\overline{\mathcal{F}}^{(p)}_{\bullet}\big{|},\big{|}\overline{\mathcal{F}}^{(p-1)}_{\bullet}\big{|};\mathbb{Q}\big{)}

is a cycle, and the resulting map of rational vector spaces

Cp\displaystyle C_{p} Hp(|¯(p)|,|¯(p1)|;)\displaystyle\to H_{p}\big{(}\big{|}\overline{\mathcal{F}}^{(p)}_{\bullet}\big{|},\big{|}\overline{\mathcal{F}}^{(p-1)}_{\bullet}\big{|};\mathbb{Q}\big{)}
[G,ω]\displaystyle[G,\omega] [fG,ω(ιn)]\displaystyle\mapsto\big{[}f^{G,\omega}_{*}(\iota_{n})\big{]}

is an isomorphism.

Proof sketch.

This follows from the description of |¯(p)|/|¯(p1)|\big{|}\overline{\mathcal{F}}^{(p)}_{\bullet}\big{|}/\big{|}\overline{\mathcal{F}}^{(p-1)}_{\bullet}\big{|} in Corollary 5.13, using the fact that [ιn]Hn((Δ1)n/H,(Δ1)n/H;)[\iota_{n}]\in H_{n}((\Delta^{1})^{n}/H,\partial(\Delta^{1})^{n}/H;\mathbb{Q}) is a generator, for any subgroup HAnH\leq A_{n}. ∎

We have not yet defined the boundary operator :CpCp1\partial\colon C_{p}\to C_{p-1}, but let us do that now. It is the difference =cd\partial=\partial_{c}-\partial_{d} between two operators, one which deletes an edge in all possible ways, and one which contracts an edge in all possible ways. Precisely, for GG a graph with p+1p+1 edges and ω=e0<<ep\omega=e_{0}<\cdots<e_{p} a total ordering on E(G)E(G), define

d[G,ω]=i=0p(1)i[G\ei,ω|E(G){ei}], and c[G,ω]=i=0p(1)i[G/ei,ω|E(G){ei}],\partial_{d}[G,\omega]=\sum_{i=0}^{p}(-1)^{i}\big{[}G\backslash e_{i},\omega|_{E(G)\setminus\{e_{i}\}}\big{]},\quad\mbox{ and }\quad\partial_{c}[G,\omega]=\sum_{i=0}^{p}(-1)^{i}\big{[}G/e_{i},\omega|_{E(G)\setminus\{e_{i}\}}\big{]},

where ω|E(G){ei}\omega|_{E(G)\setminus\{e_{i}\}} is the total order induced from ω\omega by omitting eie_{i}.

Lemma 5.17.

The map

C\displaystyle C Csing(MGr;)\displaystyle\to C_{*}^{\mathrm{sing}}(M_{\mathrm{Gr}};\mathbb{Q})
[G,ω]\displaystyle[G,\omega] fG,ω(ιn)\displaystyle\mapsto f^{G,\omega}_{*}(\iota_{n})

is a chain map.

Proof.

The boundary of (Δ1)n(\Delta^{1})^{n} is the union of the 2n2n many embeddings

(72) (Δ1)n1(Δ1)j×Δ0×(Δ1)nj11×di×1(Δ1)n(\Delta^{1})^{n-1}\xleftarrow{\approx}(\Delta^{1})^{j}\times\Delta^{0}\times(\Delta^{1})^{n-j-1}\xrightarrow{1\times d^{i}\times 1}(\Delta^{1})^{n}

for i{0,1}i\in\{0,1\} and j{0,,n1}j\in\{0,\dots,n-1\}. Writing i,j:(Δ1)n1(Δ1)n\partial_{i,j}\colon(\Delta^{1})^{n-1}\to(\Delta^{1})^{n} for this embedding and writing ιnCnsing((Δ1)n;)\iota_{n}\in C_{n}^{\mathrm{sing}}((\Delta^{1})^{n};\mathbb{Q}) for the fundamental cycle (70), the Leibniz rule for cross product leads to the convenient formula

(73) ιn\displaystyle\partial\iota_{n} =i=1n(1)i1ιi1×(ι1)×ιniCn1((Δ1)n;)\displaystyle=\sum_{i=1}^{n}(-1)^{i-1}\iota_{i-1}\times(\partial\iota_{1})\times\iota_{n-i}\in C_{n-1}(\partial(\Delta^{1})^{n};\mathbb{Z})
=i=1n(1)i1(0,j)ιn1i=1n(1)i1(1,j)ιn1.\displaystyle=\sum_{i=1}^{n}(-1)^{i-1}(\partial_{0,j})_{*}\iota_{n-1}-\sum_{i=1}^{n}(-1)^{i-1}(\partial_{1,j})_{*}\iota_{n-1}.

The lemma then follows because the composition

(Δ1)n1i,j(Δ1)nfG,ωMGr(\Delta^{1})^{n-1}\xrightarrow{\partial_{i,j}}(\Delta^{1})^{n}\xrightarrow{f^{G,\omega}}M_{\mathrm{Gr}}

is fG,ωf^{G^{\prime},\omega^{\prime}}, where GG^{\prime} is the graph obtained from GG by either collapsing (for i=0i=0) or deleting (for i=1i=1) the jjth edge in GG, and ω\omega^{\prime} is the restriction of the total order ω\omega to E(G)=E(G){ej}E(G^{\prime})=E(G)\setminus\{e_{j}\}. ∎

In order to obtain the graph spectral sequence, we must consider MGrM_{\mathrm{Gr}} with its filtration by the subspaces

FnMGr=|Fn¯|MGrF_{n}M_{\mathrm{Gr}}=|F_{n}\overline{\mathcal{F}}_{\bullet}|\subset M_{\mathrm{Gr}}

where Fn¯pF_{n}\overline{\mathcal{F}}_{p} corresponds to isomorphism classes of triangular diagrams in which b1(G0,p)nb_{1}(G_{0,p})\leq n. Fortunately the entire argument for the quasi-isomorphism C=(C,)Csing(MGr;)C=(C_{*},\partial)\simeq C_{*}^{\mathrm{sing}}(M_{\mathrm{Gr}};\mathbb{Q}) applies essentially verbatim to the simplicial subset Fn¯¯F_{n}\overline{\mathcal{F}}_{\bullet}\subset\overline{\mathcal{F}}_{\bullet}. We state the conclusion.

Theorem 5.18.

Let FnCCF_{n}C\subset C be the span of those generators [G,ω][G,\omega] for which b1(G)nb_{1}(G)\leq n. This defines a subcomplex for all nn, and the formula (71) restricts to quasi-isomorphisms

FnCCsing(FnMGr;)F_{n}C\xrightarrow{\simeq}C_{*}^{\mathrm{sing}}(F_{n}M_{\mathrm{Gr}};\mathbb{Q})

for all nn. In other words, formula (71) defines a filtered quasi-isomorphism CCsing(MGr;)C\to C_{*}^{\mathrm{sing}}(M_{\mathrm{Gr}};\mathbb{Q}).∎

By this theorem, the graph spectral sequence is isomorphic to the spectral sequence associated to the filtered chain complex (C,F)(C,F) starting on E1E^{1}, at least as a spectral sequence of rational vector spaces. It is rather easy to upgrade this isomorphism to one of spectral sequences of algebras, where (C,F)(C,F) is upgraded to a filtered DGA by declaring [G,ω][G,ω]=[GG,ωω][G,\omega][G^{\prime},\omega^{\prime}]=[G\sqcup G^{\prime},\omega\ast\omega^{\prime}], where ωω\omega\ast\omega^{\prime} denotes the “concatenation” order of E(GG)=E(G)E(G)E(G\sqcup G^{\prime})=E(G)\sqcup E(G^{\prime}) where all edges of GG come before all edges of GG^{\prime}. The coproduct is rather more involved.

Theorem 5.19.

Let C=(C,)C=(C_{*},\partial) have the coproduct induced by

(74) Δ([G,ω])=γE(G)±[G|γ,ω|γ][G/γ,ω|E(G)γ]\Delta([G,\omega])=\sum_{\gamma\subset E(G)}\pm[G_{|\gamma},\omega_{|\gamma}]\otimes[G/\gamma,\omega_{|E(G)\setminus\gamma}]

where V(G|γ)=V(G)V(G_{|\gamma})=V(G) and E(G|γ)=γE(G_{|\gamma})=\gamma while G/γG/\gamma is the graph obtained by collapsing each element of γ\gamma to a point. The edge sets of G|γG_{|\gamma} and G/γG/\gamma are in canonical bijection with γ\gamma and E(G)γE(G)\setminus\gamma respectively, and ω|γ\omega_{|\gamma} and ω|E(G)γ\omega_{|E(G)\setminus\gamma} denote the restrictions of the total order ω\omega. Finally, the sign ±\pm in each term is the sign of the shuffle SγSE(G)S_{\gamma}\in S_{E(G)} which shuffles the elements of γE(G)\gamma\subset E(G) before the elements of E(G)γE(G)\setminus\gamma.

  1. (1)

    This coproduct makes (C,F)(C,F) into a filtered coalgebra.

  2. (2)

    The map (71) is compatible with the map induced by the space level coproduct MGr=|¯||es¯||¯|×|¯|=MGr×MGrM_{\mathrm{Gr}}=|\overline{\mathcal{F}}_{\bullet}|\approx|\mathrm{es}\overline{\mathcal{F}}_{\bullet}|\to|\overline{\mathcal{F}}_{\bullet}|\times|\overline{\mathcal{F}}_{\bullet}|=M_{\mathrm{Gr}}\times M_{\mathrm{Gr}}: the diagram

    (75) C{C}C(MGr;){C_{*}(M_{\mathrm{Gr}};\mathbb{Q})}CC{C\otimes C}C(MGr×MGr;),{C_{*}(M_{\mathrm{Gr}}\times M_{\mathrm{Gr}};\mathbb{Q}),}Δ\scriptstyle{\Delta}

    commutes up to a filtration-preserving chain homotopy, where the lower horizontal map is the composition of (71)\otimes(71) and the cross product.

Proof sketch.

It is easy to see that b1(G|γ)+b1(G/γ)=b1(G)b_{1}(G_{|\gamma})+b_{1}(G/\gamma)=b_{1}(G).

The second statement is more involved, and the diagram (75) does not commute on the nose. The starting point for the homotopy is the chain ι1=ι1+cC1(Δ1;)\iota_{1}^{\prime}=\iota_{1}+\partial c\in C_{1}(\Delta^{1};\mathbb{Z}) where cC2(Δ1;)c\in C_{2}(\Delta^{1};\mathbb{Z}) is the simplex

Δ2\displaystyle\Delta^{2} Δ1\displaystyle\to\Delta^{1}
(t0,t1,t2)\displaystyle(t_{0},t_{1},t_{2}) t0(1,0)+t1(12,12)+t2(0,1).\displaystyle\mapsto t_{0}(1,0)+t_{1}\big{(}\tfrac{1}{2},\tfrac{1}{2}\big{)}+t_{2}(0,1).

In these formulas we use barycentric coordinates (t0,,tn)Δn(t_{0},\dots,t_{n})\in\Delta^{n}, in which the simplex is defined by ti0t_{i}\geq 0 and ti=1\sum t_{i}=1. In other words, ι1\iota_{1}^{\prime} is the sum (in the abelian group C1(Δ1;)C_{1}(\Delta^{1};\mathbb{Z})) of the two maps

Δ1\displaystyle\Delta^{1} Δ1\displaystyle\to\Delta^{1}
(t0,t1)\displaystyle(t_{0},t_{1}) (12t0,112t0)\displaystyle\mapsto\big{(}\tfrac{1}{2}t_{0},1-\tfrac{1}{2}t_{0}\big{)}
(t0,t1)\displaystyle(t_{0},t_{1}) (112t1,12t1),\displaystyle\mapsto\big{(}1-\tfrac{1}{2}t_{1},\tfrac{1}{2}t_{1}\big{)},

precisely the (geometric realization of the) two non-degenerate 1-simplices in the edgewise subdivision of Δ1\Delta^{1}_{\bullet}. It follows that the composition

C1(|Δ1|)C1(|es(Δ1)|)C1(|Δ1|×|Δ1|)C_{1}(|\Delta^{1}_{\bullet}|)\xrightarrow{\cong}C_{1}(|\mathrm{es}(\Delta^{1}_{\bullet})|)\to C_{1}(|\Delta^{1}_{\bullet}|\times|\Delta^{1}_{\bullet}|)

sends ι1d1×ι1+ι1×d0\iota_{1}^{\prime}\mapsto d^{1}\times\iota_{1}+\iota_{1}\times d^{0}.

Then

ιn=ι1××ι1Cn((Δ1)n;)\iota_{n}^{\prime}=\iota^{\prime}_{1}\times\dots\times\iota^{\prime}_{1}\in C_{n}((\Delta^{1})^{n};\mathbb{Z})

gives a different set of chain level representatives of the fundamental class of the cube (Δ1)n(\Delta^{1})^{n}, with the exact same formal properties as the ιn\iota_{n}. This includes the formula (73), which holds with ιn\iota^{\prime}_{n} and ιn1\iota^{\prime}_{n-1} in place of ιn\iota_{n} and ιn1\iota_{n-1}.

The formula ιn=(c)×ιn1+ι×ιn1\iota_{n}^{\prime}=(\partial c)\times\iota_{n-1}^{\prime}+\iota\times\iota_{n-1}^{\prime} implies inductively that ιn=ιn+cn\iota^{\prime}_{n}=\iota_{n}+\partial c_{n} for

cn=j=1nιj1×c×ιnj.c_{n}=\sum_{j=1}^{n}\iota_{j-1}\times c\times\iota^{\prime}_{n-j}.

Then the two chain maps

C\displaystyle C Csing(MGr;)\displaystyle\to C_{*}^{\mathrm{sing}}(M_{\mathrm{Gr}};\mathbb{Q})
[G,ω]\displaystyle[G,\omega] fG,ω(ιn)\displaystyle\mapsto f^{G,\omega}_{*}(\iota_{n})
[G,ω]\displaystyle[G,\omega] fG,ω(ιn)\displaystyle\mapsto f^{G,\omega}_{*}(\iota^{\prime}_{n})

are chain homotopic via a filtration-preserving chain homotopy. One then checks that the diagram (75) commutes strictly, if the top horizontal map is replaced by the chain homotopic map [G,ω]fG,ω(ιn)[G,\omega]\mapsto f^{G,\omega}_{*}(\iota^{\prime}_{n}).

To see this, we study the space-level coproduct ΔX:|X||es(X)||X|×|X|\Delta_{X}\colon|X|\approx|\mathrm{es}(X)|\to|X|\times|X| from Definition 3.5 and especially the induced chain map chain map

Cn(|Δ1|E)(Δ(Δ1)E)Cn(|Δ1|E×|Δ1|E)C_{n}(|\Delta^{1}_{\bullet}|^{E})\xrightarrow{(\Delta_{(\Delta^{1}_{\bullet})^{E}})_{*}}C_{n}(|\Delta^{1}_{\bullet}|^{E}\times|\Delta^{1}_{\bullet}|^{E})

in the case X=(Δ1)EX=(\Delta^{1}_{\bullet})^{E} for a finite set EE. Given also a bijection ω:{0,,n1}E\omega\colon\{0,\dots,n-1\}\to E we have defined the fundamental chains ιn\iota_{n} and ιn\iota^{\prime}_{n} in the domain of this chain map, and claim the formula

(76) (Δ(Δ1)E)(ιn)=γE±(|Lγ|,|Rγ|)ιn,(\Delta_{(\Delta^{1}_{\bullet})^{E}})_{*}(\iota^{\prime}_{n})=\sum_{\gamma\subset E}\pm(|L_{\gamma}|,|R_{\gamma}|)_{*}\iota_{n},

where Lγ=eELγ,eL_{\gamma}=\prod_{e\in E}L_{\gamma,e} and Rγ=eERγ,eR_{\gamma}=\prod_{e\in E}R_{\gamma,e} for maps of simplicial sets Lγ,e,Rγ,e:Δ1Δ1L_{\gamma,e},R_{\gamma,e}\colon\Delta^{1}_{\bullet}\to\Delta^{1}_{\bullet} given as

Lγ,e\displaystyle L_{\gamma,e} ={idif eγd1s0if eγ\displaystyle=\begin{cases}\mathrm{id}&\text{if $e\in\gamma$}\\ d^{1}s^{0}&\text{if $e\not\in\gamma$}\end{cases} Rγ,e={d0s0if eγidif eγ.\displaystyle R_{\gamma,e}=\begin{cases}d^{0}s^{0}&\text{if $e\in\gamma$}\\ \mathrm{id}&\text{if $e\not\in\gamma$}.\end{cases}

We have already seen formula (76) in the case n=1n=1, and the general case follows by expanding the nn-fold cross product (d1×ι1+ι1×d0)×nCn((Δ1×Δ1)n)(d^{1}\times\iota_{1}+\iota_{1}\times d^{0})^{\times n}\in C_{n}((\Delta^{1}\times\Delta^{1})^{n}) as a sum of 2n2^{n} terms and using graded-commutativity of the cross product (i.e., that ×:C(X)C(Y)C(X×Y)\times\colon C_{*}(X)\otimes C_{*}(Y)\to C_{*}(X\times Y) turns singular chains into a lax symmetric monoidal functor from spaces to chain complexes, in the usual symmetric monoidal structure on chain complexes). The sign arising when permuting factors is precisely the sign of the shuffle SγSE(G)S_{\gamma}\in S_{E(G)} mentioned in the statement of the theorem.

For each γE\gamma\subset E, the map (Lγ,Rγ):(Δ1)E(Δ1)E×(Δ1)E(L_{\gamma},R_{\gamma})\colon(\Delta^{1})^{E}\to(\Delta^{1})^{E}\times(\Delta^{1})^{E} has 2n2n coordinates, nn of which are the projections (Δ1)EΔ1(\Delta^{1})^{E}\to\Delta^{1} and the remaining nn coordinates are constant (either d0d^{0} or d1d^{1}). Setting E=E(G)E=E(G), we have canonical maps fG:(Δ1)EMGrf^{G}\colon(\Delta^{1})^{E}\to M_{\mathrm{Gr}} as in Definition 5.15. Using, as in the proof of Lemma 5.17, that precomposing with d1:Δ0Δ1d^{1}\colon\Delta^{0}\to\Delta^{1} in one of the factors of (Δ1)E(\Delta^{1})^{E} corresponds to deleting the corresponding edge, we see that fGLγ=fG|γπγf^{G}\circ L_{\gamma}=f^{G|\gamma}\circ\pi_{\gamma}, where πγ:(Δ1)E(Δ1)γ\pi_{\gamma}:(\Delta^{1})^{E}\to(\Delta^{1})^{\gamma} denotes the projection, and similarly fGRγ=fG/γπEγf^{G}\circ R_{\gamma}=f^{G/\gamma}\circ\pi_{E\setminus\gamma}, using that precomposing with d0d^{0} corresponds to collapsing an edge. We can summarize this situation by a commutative diagram of spaces

(Δ1)E{(\Delta^{1})^{E}}(Δ1)γ×(Δ1)Eγ{(\Delta^{1})^{\gamma}\times(\Delta^{1})^{E\setminus\gamma}}MGr×MGr{{M_{\mathrm{Gr}}\times M_{\mathrm{Gr}}}}(Δ1)E{(\Delta^{1})^{E}}(Δ1)E×(Δ1)E{(\Delta^{1})^{E}\times(\Delta^{1})^{E}}MGr×MGr.{{M_{\mathrm{Gr}}\times M_{\mathrm{Gr}}}.}(πγ,πEγ)\scriptstyle{(\pi_{\gamma},\pi_{E\setminus\gamma})}\scriptstyle{\approx}(fG|γ,fG/γ)\scriptstyle{(f^{G|\gamma},f^{G/\gamma})}(Lγ,Rγ)\scriptstyle{(L_{\gamma},R_{\gamma})}(fG,fG)\scriptstyle{(f^{G},f^{G})}

Using the same notations ιn,ιnCn((Δ1)E)\iota_{n},\iota^{\prime}_{n}\in C_{n}((\Delta^{1})^{E}) for the images of ιn,ιnCn((Δ1)n)\iota_{n},\iota^{\prime}_{n}\in C_{n}((\Delta^{1})^{n}) under the homeomorphism (Δ1)n(Δ1)E(\Delta^{1})^{n}\approx(\Delta^{1})^{E} induced by ω\omega, and similarly ι|γ|C((Δ1)γ)\iota_{|\gamma|}\in C_{*}((\Delta^{1})^{\gamma}) and ιn|γ|C((Δ1)Eγ)\iota_{n-|\gamma|}\in C_{*}((\Delta^{1})^{E\setminus\gamma}) using the restriction of the total order to the subsets γE\gamma\subset E and EγEE\setminus\gamma\subset E, naturality of ΔX:|X||X|×|X|\Delta_{X}\colon|X|\to|X|\times|X| then shows that

(Δ¯)fG(ιn)=γE(G)±(fG,fG)(Lγ,Rγ)ιn\displaystyle(\Delta_{\overline{\mathcal{F}}_{\bullet}})_{*}f^{G}_{*}(\iota^{\prime}_{n})=\sum_{\gamma\subset E(G)}\pm(f^{G},f^{G})_{*}(L_{\gamma},R_{\gamma})_{*}\iota_{n} =γE(G)±(fG|γ,fG/γ)(ι|γ|×ιn|γ|)\displaystyle=\sum_{\gamma\subset E(G)}\pm(f^{G|\gamma},f^{G/\gamma})_{*}(\iota_{|\gamma|}\times\iota_{n-|\gamma|})
=γE(G)±(fG|γι|γ|)×(fG/γιn|γ|).\displaystyle=\sum_{\gamma\subset E(G)}\pm(f^{G|\gamma}_{*}\iota_{|\gamma|})\times(f^{G/\gamma}_{*}\iota_{n-|\gamma|}).\qed
Corollary 5.20.

Let Gr(C)\operatorname{Gr}(C) denote the associated graded of the chain complex C=(C,)C=(C_{*},\partial), filtered by the first Betti numbers of the graphs, with bigrading in which a generator [G,ω][G,\omega] has bidegree (b1(G),|E(G)|b1(G))(b_{1}(G),|E(G)|-b_{1}(G)). The boundary map \partial has bidegree (0,1)(0,-1).

The map (71) induces an isomorphism of bigraded bialgebras H(Gr(C))E,1GH_{*}(\operatorname{Gr}(C))\to{}^{G}\!E^{1}_{*,*} where the coproduct on H(Gr(C))H_{*}(\operatorname{Gr}(C)) is induced by the Connes–Kreimer formula (74).

Each generator [G,ω]C[G,\omega]\in C represents a non-zero element of Grg(C)\operatorname{Gr}_{g}(C) for g=b1(G)g=b_{1}(G) which we denote by the same notation [G,ω][G,\omega]. Then Gr(C)\operatorname{Gr}(C) inherits a product and coproduct from CC, making it a bigraded bialgebra (isomorphic to CC as a bigraded bialgebra). The boundary map =cd:CC\partial=\partial_{c}-\partial_{d}\colon C\to C induces a boundary map on Gr(C)\operatorname{Gr}(C) given by a similar formula on generators [G,ω]Gr(C)[G,\omega]\in\operatorname{Gr}(C) but omitting the terms involving [G,ω][G^{\prime},\omega^{\prime}] where b1(G)<b1(G)b_{1}(G^{\prime})<b_{1}(G). Precisely, this amounts to dropping the terms in c([G,ω])\partial_{c}([G,\omega]) in which a loop is collapsed, and omitting the terms in d([G,ω])\partial_{d}([G,\omega]) in which a non-bridge is deleted.

Proof.

We have already shown that (71) is a filtered quasi-isomorphism, so it induces an isomorphism on homology. The filtered homotopy in the theorem implies that the isomorphism on homology is a map of coalgebras. That it is also a map of algebras is similar but easier, we omit the details. ∎

5.5. Comparison with the full graph complex

In the chain complex Gr(C)\operatorname{Gr}(C), some remnant of d\partial_{d} will remain in the boundary map, namely a (signed) sum of bridge deletion. Here, a bridge in a not necessarily connected graph is any edge whose deletion increases the number of connected components. Note that deleting a bridge results in a graph that is not connected, and the result is typically zero in Indec(Gr(C))\mathrm{Indec}(\operatorname{Gr}(C)).

Motivated by this observation, we pass to indecomposables of Gr(C)\operatorname{Gr}(C) to establish a precise relationship between the E1E^{1}-page of the graph spectral sequence and the cochain complex 𝖦𝖢2\mathsf{GC}_{2} considered by Kontsevich, Willwacher, and others, in which all graphs are connected and the differential involves only edge contraction, not deletion. We refer to [Wil15, §3] for the precise definition of 𝖦𝖢2\mathsf{GC}_{2} and its variants considered below. However, from now on we take the following bigrading on 𝖦𝖢2\mathsf{GC}_{2} and all of its variants, which differs from the grading used by Willwacher: the generators in bidegree (s,t)(s,t) are graphs with first Betti number ss and s+ts+t edges. (This graph complex, with this bigrading, is called 𝖦𝖢0\mathsf{GC}_{0} in some other sources, such as [Bro21b, Wil15], but we will stick with the notation 𝖦𝖢2\mathsf{GC}_{2}.)

Recall the full graph complex denoted 𝖿𝖦𝖢2\mathsf{fGC}_{2}^{\circlearrowleft} by Willwacher in [Wil15, §3] and denoted 𝖿𝖦𝖢2\mathsf{fGC}_{2} in [KWŽ17]. Like CC^{\vee}, it is also built out of all graphs with no conditions on connectivity or the valency of vertices, and where “tadpoles” are allowed. Reading the definition closely though, one notices that the graphs in 𝖿𝖦𝖢2\mathsf{fGC}_{2}^{\circlearrowleft} must be non-empty, which ours need not be. By comparing definitions we therefore obtain isomorphisms of rational vector spaces

(77) (C/.1)𝖿𝖦𝖢2,C𝖿𝖦𝖢2.,(C/\mathbb{Q}.1)^{\vee}\cong\mathsf{fGC}_{2}^{\circlearrowleft},\quad\quad C^{\vee}\cong\mathsf{fGC}_{2}^{\circlearrowleft}\oplus\mathbb{Q}.\emptyset^{\vee},

arising from bijections between basis vectors, where \emptyset denotes the graph with no vertices and no edges, representing the unit 1C1\in C. It is of some importance for us to include the empty graph, in order for our bialgebra to be unital and co-unital. Geometrically, (C/.1)(C/\mathbb{Q}.1)^{\vee} can be interpreted as the relative cellular cochain complex of the pair (MGr,{})(M_{\mathrm{Gr}},\{\emptyset\}), where {}MGr\{\emptyset\}\subset M_{\mathrm{Gr}} is the path component corresponding to the empty graph, as in Corollary 5.27.

The usual differential δ\delta on 𝖿𝖦𝖢2\mathsf{fGC}_{2}^{\circlearrowleft}, defined in [Wil15, Remark 3.3] by “inserting an edge in all possible ways,” does not correspond to :CC\partial^{\vee}\colon C^{\vee}\to C^{\vee}, though: recall that =cd\partial=\partial_{c}-\partial_{d} where c\partial_{c} is defined as alternating sum of edge contraction while d\partial_{d} is defined as alternating sum of edge deletion, and most of the terms in d\partial_{d} are not seen in the dual formula for δ\delta on 𝖿𝖦𝖢2\mathsf{fGC}^{\circlearrowleft}_{2}.

Passing to indecomposables of the bialgebra CC corresponds to passing to the subcomplex 𝖿𝖦𝖢2,conn𝖿𝖦𝖢2\mathsf{fGC}^{\circlearrowleft}_{2,\mathrm{conn}}\subset\mathsf{fGC}^{\circlearrowleft}_{2} from [Wil15, §3] spanned by connected graphs. Indeed, Indec(C)=K/K2\mathrm{Indec}(C)=K/K^{2} where KK is the kernel of the augmentation map CC\to\mathbb{Q}; here, \mathbb{Q} is concentrated in homological degree 0, and the augmentation sends [G]1[G]\mapsto 1 for any graph GG with E(G)=E(G)=\emptyset. In particular, any graph with edges is in KK. Let [p]C[p]\in C denote the class of a graph consisting of a single vertex and no edges. We next record, for convenience, a vector space basis for K2K^{2}: it consists of

  1. (1)

    [G,<][G,<] where GG ranges over graphs that are disjoint unions of two graphs with edges, one per isomorphism class;

  2. (2)

    [G,<]f([p])[G,<]\cdot f([p]), where GG is any alternating, connected graph with at least one edge, one per isomorphism class, and f[t]f\in\mathbb{Q}[t] ranges over a basis for polynomials with f(1)=0f(1)=0;

  3. (3)

    g([p])g([p]) where g[t]g\in\mathbb{Q}[t] ranges over a basis for polynomials with g(1)=g(1)=0g(1)=g^{\prime}(1)=0.

Here a graph is called alternating if its automorphism group acts on its edge set with only permutations of positive sign. Then K/K2K/K^{2} has a basis given by classes of elements [G,<][G,<] as GG ranges over alternating, connected graphs with at least one edge, one per isomorphism class, together with the class of [p]1K[p]-1\in K. Removing the class of [p]1[p]-1 from this set, and adding the classes of both [p][p] and 11, gives a basis of C/K2C/K^{2}. Therefore the composition of the natural maps

(78) K/K2C/K2C/(K2+.1)K/K^{2}\hookrightarrow C/K^{2}\twoheadrightarrow C/(K^{2}+\mathbb{Q}.1)

is an isomorphism of chain complexes. Now CC has a basis as a rational vector space given by classes [G,<][G,<] where we choose one graph GG for each isomorphism class, and the subset given by connected (in particular non-empty) graphs maps to a basis of C/(K2+.1)C/(K^{2}+\mathbb{Q}.1). The isomorphism (77) therefore restricts to an isomorphism (C/(K2+.1))𝖿𝖦𝖢2,conn(C/(K^{2}+\mathbb{Q}.1))^{\vee}\cong\mathsf{fGC}^{\circlearrowleft}_{2,\mathrm{conn}}, which we combine with (78) to obtain an isomorphism of vector spaces

𝖿𝖦𝖢2,connIndec(C).\mathsf{fGC}^{\circlearrowleft}_{2,\mathrm{conn}}\xrightarrow{\cong}\mathrm{Indec}(C)^{\vee}.

We emphasize that the graded vector space K/K2K/K^{2} is 1-dimensional in homological degree 0, spanned by the class of [p]1K[p]-1\in K. The dual basis element is denoted ([p]1)([p]-1)^{\vee} and corresponds to [p]𝖿𝖦𝖢2,conn[p]^{\vee}\in\mathsf{fGC}^{\circlearrowleft}_{2,\mathrm{conn}} under the isomorphism induced by (78). As before, the differential δ\delta on the left hand side of this isomorphism does not correspond to the full =cd:CC\partial^{\vee}=\partial_{c}^{\vee}-\partial_{d}^{\vee}\colon C^{\vee}\to C^{\vee}. This discrepancy goes away when replacing CC with Gr(C)\operatorname{Gr}(C) though, and we have the following precise statement.

Lemma 5.21.

The analogue of the map (78) for the differential graded Hopf algebra Gr(C)\operatorname{Gr}(C) induces an isomorphism of differential graded Lie algebras

(79) 𝖿𝖦𝖢2,connIndec(Gr(C)).\mathsf{fGC}^{\circlearrowleft}_{2,\mathrm{conn}}\xrightarrow{\cong}\mathrm{Indec}(\operatorname{Gr}(C))^{\vee}.
Proof.

CC and Gr(C)\operatorname{Gr}(C) are canonically isomorphic as graded bialgebras, using the basis given by the [G,<][G,<] which is canonical up to signs. Passing from CC to Gr(C)\operatorname{Gr}(C) has the effect of omitting those terms of ([G,<])\partial([G,<]) involving graphs GG^{\prime} with b1(G)<b1(G)b_{1}(G^{\prime})<b_{1}(G), which amounts to omitting those terms of d([G,<])\partial_{d}([G,<]) in which a non-bridge is deleted and those terms of c([G,<])\partial_{c}([G,<]) in which a tadpole is collapsed. We now inspect what happens to the boundary homomorphism when passing from Gr(C)\operatorname{Gr}(C) to Indec(Gr(C))=K/K2\mathrm{Indec}(\operatorname{Gr}(C))=K/K^{2} where K=Ker(Gr(C))K=\mathrm{Ker}(\operatorname{Gr}(C)\to\mathbb{Q}), starting with a brief analysis of the relations in K2K^{2}. If γ=γγ′′\gamma=\gamma^{\prime}\sqcup\gamma^{\prime\prime} where E(γ)E(γ′′)E(\gamma^{\prime})\neq\emptyset\neq E(\gamma^{\prime\prime}), then [γ,<]K2[\gamma,<]\in K^{2}. Therefore [γ,<]0[\gamma,<]\equiv 0 when γ\gamma has two edges in distinct components of γ\gamma. In contrast, if γ′′=p\gamma^{\prime\prime}=p consists of a single vertex and no edges and E(γ)E(\gamma^{\prime})\neq\emptyset, then [p][γ,<][γ,<]K2[p][\gamma^{\prime},<]-[\gamma^{\prime},<]\in K^{2}. In other words, modulo K2K^{2} we have the relation [pγ,<][γ,<][p\sqcup\gamma^{\prime},<]\equiv[\gamma^{\prime},<] which allows us to remove isolated vertices from any graph with at least one edge.

Modulo K2Gr(C)K^{2}\subset\operatorname{Gr}(C), most terms of d([G,<])\partial_{d}([G,<]) vanish, because deleting a bridge results in a disconnected graph. The exception is “bridges to nowhere”: if v0V(G)v_{0}\in V(G) has valence 1 then it is connected to the rest of GG by a single edge e0E(G)e_{0}\in E(G) and there is a term in d([G,<])\partial_{d}([G,<]) of the form ±[{v0}G,<]=±[p][G,<]\pm[\{v_{0}\}\sqcup G^{\prime},<^{\prime}]=\pm[p][G^{\prime},<^{\prime}] where GGG^{\prime}\subset G is the subgraph with V(G)=V(G){v0}V(G^{\prime})=V(G)\setminus\{v_{0}\} and E(G)=E(G){e0}E(G^{\prime})=E(G)\setminus\{e_{0}\}. Assuming E(G)E(G^{\prime})\neq\emptyset, we have [p][(G,<)][(G,<)][p][(G^{\prime},<^{\prime})]\equiv[(G^{\prime},<^{\prime})] modulo K2K^{2} and this term does not vanish. This analysis shows that d\partial_{d} induces a homomorphism Indec(Gr(C))nIndec(Gr(C))n1\mathrm{Indec}(\operatorname{Gr}(C))_{n}\to\mathrm{Indec}(\operatorname{Gr}(C))_{n-1} for n2n\geq 2 given as alternating sum of deleting such bridges-to-nowhere together with their 1-valent vertex. These terms in d([G,<])\partial_{d}([G,<]) precisely match with those terms in c([G,<])\partial_{c}([G,<]) in which the corresponding edge is collapsed, and we deduce that the boundary homomorphism on Indec(Gr(C))\mathrm{Indec}(\operatorname{Gr}(C)) induced by =cd\partial=\partial_{c}-\partial_{d} is given on representatives [G,<][G,<] with GG connected and n=|E(G)|2n=|E(G)|\geq 2 by the formula

[G,<]=e(1)ω(e)[G/e,<|E(G){e}],\partial[G,<]=\sum_{e}(-1)^{\omega(e)}[G/e,<_{|E(G)\setminus\{e\}}],

where eE(G)e\in E(G) runs over all edges except edges that are bridges-to-nowhere or tadpoles, and ω:E(G){0,1,,n1}\omega\colon E(G)\to\{0,1,\dots,n-1\} is an order-preserving bijection. In the special case n=1n=1 we get ([G,<])=[p][p]21[p]\partial([G,<])=[p]-[p]^{2}\equiv 1-[p] modulo K2K^{2} when GG consists of two vertices connected by an edge, and ([G,<])=0\partial([G,<])=0 when GG is a tadpole. Altogether, dualizing this homomorphism \partial precisely matches with the differential δ:𝖿𝖦𝖢2,conn𝖿𝖦𝖢2,conn\delta\colon\mathsf{fGC}_{2,\mathrm{conn}}^{\circlearrowleft}\to\mathsf{fGC}_{2,\mathrm{conn}}^{\circlearrowleft}, see [Wil15, Remark 3.3] and [Wil15, Figure 3].

The Lie cobracket on Indec(Gr(C))=K/K2\mathrm{Indec}(\operatorname{Gr}(C))=K/K^{2} is obtained by anti-symmetrizing the Connes–Kreimer formula, restricting to K=Ker(Gr(C))K=\mathrm{Ker}(\operatorname{Gr}(C)\to\mathbb{Q}) and reducing modulo K2K^{2}. We shall first consider the terms of (74) which only involve graphs with at least one edge. If GG is a connected graph with at least one edge, then the sum (74) will likely contain terms in which γG\gamma\subset G is not connected, and such terms must be reduced modulo K2K^{2}. Since GG is connected, reducing the right-hand side of (74) modulo K2Gr(C)K^{2}\otimes\operatorname{Gr}(C), and restricting attention to those terms in which γ\gamma has at least one edge, lets us remove isolated vertices from γ\gamma and lets us cancel terms in which γ\gamma contains two edges in distinct components of γ\gamma. After anti-symmetrizing and dualizing, the sum of these terms, in which γE(G)\emptyset\subsetneq\gamma\subsetneq E(G), gives a formula for the Lie bracket between any two elements of Indec(Gr(C))\mathrm{Indec}(\operatorname{Gr}(C))^{\vee} of positive cohomological degree, and the formula matches with the Lie bracket on 𝖿𝖦𝖢2,conn\mathsf{fGC}^{\circlearrowleft}_{2,\mathrm{conn}} explained in [KWŽ17, §2] and implicit in [Wil15, §3], given by anti-symmetrized insertion.

Now we inspect the terms in the antisymmetrization of (74) with E(γ)=E(\gamma)=\emptyset or E(G/γ)=E(G/\gamma)=\emptyset. Writing y=[G,<]y=[G,<] with E(G)E(G)\neq\emptyset and v=|V(G)|v=|V(G)|, the sum of the terms of the antisymmetrized coproduct of yy that involve graphs with no edges is

([p]v[p])y+y([p][p]v).([p]^{v}-[p])\otimes y+y\otimes([p]-[p]^{v}).

Note [p]v[p](v1)([p]1)[p]^{v}-[p]\equiv(v\!-\!1)\cdot([p]-1) modulo K2K^{2}. Since [p]1K[p]-1\in K reduces to a basis for the degree-zero part of K/K2=Indec(Gr(C))K/K^{2}=\mathrm{Indec}(\operatorname{Gr}(C)), we obtain a formula for the Lie bracket with the element ([p]1)Indec(Gr(C))([p]-1)^{\vee}\in\mathrm{Indec}(\operatorname{Gr}(C))^{\vee}: for E(G)E(G)\neq\emptyset, it is given by [G,<](v1)[G,<][G,<]^{\vee}\mapsto(v-1)[G,<]^{\vee} where v=|V(G)|v=|V(G)| is the number of vertices of GG. This again agrees with the “anti-symmetrized insertion” Lie bracket with [p]𝖿𝖦𝖢2,conn[p]^{\vee}\in\mathsf{fGC}_{2,\mathrm{conn}}^{\circlearrowleft}. ∎

Proposition 5.22.

Let

L=Indec(Gr(C))L^{\vee}=\mathrm{Indec}(\operatorname{Gr}(C))

be the graph complex obtained as the indecomposables in the differential graded bialgebra Gr(C)\operatorname{Gr}(C). We have an isomorphism of bigraded Lie coalgebras

Indec(E,1G)H(L)\mathrm{Indec}\big{(}{}^{G}\!E^{1}_{*,*}\big{)}\cong H_{*}(L^{\vee})

and a dual isomorphism of bigraded Lie algebras

Prim(E1,G)H(L),\mathrm{Prim}\big{(}{}^{G}\!E_{1}^{*,*}\big{)}\cong H^{*}(L),

where L=Indec(Gr(C))𝖿𝖦𝖢2,connL=\mathrm{Indec}(\operatorname{Gr}(C))^{\vee}\cong\mathsf{fGC}^{\circlearrowleft}_{2,\mathrm{conn}}. Moreover, there is a split extension of Lie algebras

(80) H(𝖦𝖢2)H(L)LbivH^{*}(\mathsf{GC}_{2})\hookrightarrow H^{*}(L)\twoheadrightarrow L^{\mathrm{biv}}

where LbivL^{\mathrm{biv}} denotes the bigraded vector space which is \mathbb{Q} in bidegrees (1,4k)(1,4k) for all k0k\geq 0 and is 0 in all other bidegrees, equipped with the trivial Lie algebra structure.

Proof.

We have already constructed a filtered quasi-isomorphism between CC and the singular chains of BKGrBK_{\mathrm{Gr}}, giving an isomorphism

E,1GH(Gr(C)).{}^{G}\!E^{1}_{*,*}\cong H_{*}(\operatorname{Gr}(C)).

To get to the first statement in the proposition, we would like to use an isomorphism of the form Indec(H(A))H(Indec(A))\mathrm{Indec}(H_{*}(A))\cong H_{*}(\mathrm{Indec}(A)) when AA is the differential graded bialgebra Gr(C)\operatorname{Gr}(C). Independently of the coproduct, there is a natural transformation

Indec(H(A))H(Indec(A))\mathrm{Indec}(H_{*}(A))\to H_{*}(\mathrm{Indec}(A))

defined for any differential graded algebra AA equipped with an augmentation ϵ:A\epsilon\colon A\to\mathbb{Q}, and it is well known that this transformation is an isomorphism whenever H(A)H_{*}(A) is a free graded-commutative algebra and AA is semifree. This is proved by reducing to the case where A=0\partial_{A}=0. (Recall that a CDGA is semifree when the underlying graded-commutative algebra, obtained by forgetting the differential, is free.)

Now Gr(C)\operatorname{Gr}(C) is certainly semifree: as generators we can take [G,<][G,<] with GG connected, one such for each isomorphism class which does not possess automorphisms inducing an odd permutation of E(G)E(G). Unfortunately H(Gr(C))H_{*}(\operatorname{Gr}(C)) is not quite free, because of the relation x2=xx^{2}=x in degree 0, where x=[p]x=[p] is the class of a graph with one vertex and no edges, see Corollary 5.27. To get around this nuisance, we artificially add two algebra generators

Gr(C)A=Gr(C)[y]Λ[z],\operatorname{Gr}(C)\hookrightarrow A=\operatorname{Gr}(C)\otimes\mathbb{Q}[y]\otimes\Lambda_{\mathbb{Q}}[z],

with deg(y)=0\mathrm{deg}(y)=0 and deg(z)=1\mathrm{deg}(z)=1, and with boundary map extended as y=0\partial y=0 and z=xy1\partial z=xy-1. We do not attempt to extend the chain level coproduct, but extend the augmentation as ϵ(y)=1\epsilon(y)=1. Then the class [x]H0(Gr(C))[x]\in H_{0}(\operatorname{Gr}(C)) maps to a unit in H0(A)H_{0}(A) and the inclusion of Gr(C)\operatorname{Gr}(C) induces an isomorphism from the localized homology algebra

H(Gr(C))H(Gr(C))[x1]H(A).H_{*}(\operatorname{Gr}(C))\twoheadrightarrow H_{*}(\operatorname{Gr}(C))[x^{-1}]\xrightarrow{\cong}H_{*}(A).

Since [x]H0(Gr(C))[x]\in H_{0}(\operatorname{Gr}(C)) is a unit modulo K=Ker(H(Gr(C)))K=\mathrm{Ker}(H_{*}(\operatorname{Gr}(C))\to\mathbb{Q}), inverting xx does not change the indecomposables K/K2K/K^{2}, so we have

Indec(H(Gr(C)))Indec(H(A)).\mathrm{Indec}(H_{*}(\operatorname{Gr}(C)))\xrightarrow{\cong}\mathrm{Indec}(H_{*}(A)).

By inspection, we also see that the induced chain level map

Indec(Gr(C))Indec(A)\mathrm{Indec}(\operatorname{Gr}(C))\hookrightarrow\mathrm{Indec}(A)

is a quasi-isomorphism (on the chain level, the cokernel has two additive generators zz and y1y-1, but these are connected by a non-zero boundary map).

Inverting xx turns the ring H0(Gr(C))=[x]/(x2x)H_{0}(\operatorname{Gr}(C))=\mathbb{Q}[x]/(x^{2}-x) into [x]/(x1)\mathbb{Q}[x]/(x-1)\cong\mathbb{Q}, and the localization H(Gr(C))[x1]H_{*}(\operatorname{Gr}(C))[x^{-1}] inherits a coproduct, making it a connected commutative Hopf algebra. By a classical theorem of Leray (see [MM65, Theorem 7.5]), this implies that the underlying graded algebra is isomorphic to a free graded-commutative algebra, so we finally deduce a commutative diagram of isomorphisms

Indec(E,1G)=Indec(H(Gr(C))){\mathrm{Indec}({}^{G}\!E^{1}_{*,*})=\mathrm{Indec}(H_{*}(\operatorname{Gr}(C)))}H(Indec(Gr(C)))=H(L){H_{*}(\mathrm{Indec}(\operatorname{Gr}(C)))=H_{*}(L^{\vee})}Indec(H(A)){\mathrm{Indec}(H_{*}(A))}H(Indec(A)),{H_{*}(\mathrm{Indec}(A)),}\scriptstyle{\cong}\scriptstyle{\cong}\scriptstyle{\cong}

since AA is semifree and H(A)H_{*}(A) is free. The formula Prim(E1,G)H(L)\mathrm{Prim}({}^{G}\!E_{1}^{*,*})\cong H^{*}(L) is obtained by dualizing.

Finally, Willwacher shows [Wil15, Proposition 3.4] that the cohomology of L𝖿𝖦𝖢2,connL\cong\mathsf{fGC}^{\circlearrowleft}_{2,\mathrm{conn}} splits as a vector space as

H(L)H(𝖦𝖢2)Lbiv,H^{*}(L)\cong H^{*}(\mathsf{GC}_{2})\oplus L^{\mathrm{biv}},

where the second summand is spanned by “loop” graphs L4k+1L_{4k+1}, k0k\geq 0, with 4k+14k+1 edges connecting 4k+14k+1 bivalent vertices in a circular pattern (or rather, the dual basis vectors to such graphs, spanning a cochain complex with trivial differential, so LbivH(Lbiv)L^{\mathrm{biv}}\cong H^{*}(L^{\mathrm{biv}})). The inclusion H(𝖦𝖢2)H(L)H^{*}(\mathsf{GC}_{2})\hookrightarrow H^{*}(L) is a Lie algebra homomorphism by comparing definitions. For degree reasons, this additive splitting makes

H(𝖦𝖢2)H(L)LbivH^{*}(\mathsf{GC}_{2})\hookrightarrow H^{*}(L)\twoheadrightarrow L^{\mathrm{biv}}

into a split extension of Lie algebras, and LbivL^{\mathrm{biv}} must have trivial Lie bracket. ∎

The map of filtered spaces MGrBKGrBK()M_{\mathrm{Gr}}\simeq_{\mathbb{Q}}BK_{\mathrm{Gr}}\to BK(\mathbb{Z}) induces a map of spectral sequences of bialgebras, of the form

(81) E,rGE,rQ{}^{G}\!E^{r}_{*,*}\to{}^{Q}\!E^{r}_{*,*}

abutting to the induced map

H(MGr)H(BK()).H_{*}(M_{\mathrm{Gr}})\to H_{*}(BK(\mathbb{Z})).

When passing to primitives in the linear duals of the respective E1E^{1}-pages, it therefore induces a map of Lie algebras

Prim(E1,Q)Prim(E1,G)H(L)LbivH(𝖦𝖢2)\mathrm{Prim}\big{(}{}^{Q}\!E_{1}^{*,*}\big{)}\to\mathrm{Prim}\big{(}{}^{G}\!E_{1}^{*,*}\big{)}\cong H^{*}(L)\cong L^{\mathrm{biv}}\oplus H^{*}(\mathsf{GC}_{2})

from the primitives in the E1E_{1}-page of the (cohomological) Quillen spectral sequence to the semidirect product corresponding to the split extension (80). The bidegrees are such that primitives in E1s,tQ{}^{Q}\!E_{1}^{s,t} are mapped to graphs with first Betti number ss and s+ts+t many edges. Restricted to diagonal bidegrees, this can be identified with a map of Lie algebras

(82) Prim(gE1g,gQ)Prim(gE1g,gG)H0(𝖦𝖢2)𝔤𝔯𝔱,\mathrm{Prim}\Big{(}\bigoplus_{g}{}^{Q}\!E_{1}^{g,g}\Big{)}\to\mathrm{Prim}\Big{(}\bigoplus_{g}{}^{G}\!E_{1}^{g,g}\Big{)}\cong H^{0}(\mathsf{GC}_{2})\cong\mathfrak{grt},

with the last isomorphism following from [Wil15, Theorem 1.1]. Here 𝔤𝔯𝔱\mathfrak{grt} is the Grothendieck–Teichmüller Lie algebra (denoted 𝔤𝔯𝔱1\mathfrak{grt}_{1} in op. cit.).

5.5.1. Relationship with the double complex of [KWŽ17]

The full differential on Indec(C)\mathrm{Indec}(C)^{\vee} induced by dualizing \partial can be related to a certain double complex considered in [KWŽ17]. Let us point out a notational difference: in the paper [Wil15], Willwacher uses a superscript \circlearrowleft on graph complexes to denote that tadpoles are allowed and the lack of such a superscript means that they are disallowed, while [KWŽ17] uses a superscript to denote that tadpoles are disallowed and the lack of superscript means they are allowed. We follow the notation of the former.

Proposition 5.23.

Let :𝖿𝖦𝖢2,conn𝖿𝖦𝖢2,conn\nabla\colon\mathsf{fGC}^{\circlearrowleft}_{2,\mathrm{conn}}\to\mathsf{fGC}^{\circlearrowleft}_{2,\mathrm{conn}} be the differential given by Lie bracket with a tadpole graph, as in [KWŽ17]. Then the composition (78) induces an isomorphism of differential graded Lie algebras

(𝖿𝖦𝖢2,conn,δ+)Indec(C).(\mathsf{fGC}^{\circlearrowleft}_{2,\mathrm{conn}},\delta+\nabla)\xrightarrow{\cong}\mathrm{Indec}(C)^{\vee}.

Let us also remark that [G,<](1)|V(G)|[G,<][G,<]\mapsto(-1)^{|V(G)|}[G,<] defines an isomorphism of complexes (𝖿𝖦𝖢2,conn,δ+)(𝖿𝖦𝖢2,conn,δ)(\mathsf{fGC}^{\circlearrowleft}_{2,\mathrm{conn}},\delta+\nabla)\to(\mathsf{fGC}^{\circlearrowleft}_{2,\mathrm{conn}},\delta-\nabla), so it does not matter much if this proposition is stated with δ+\delta+\nabla or δ\delta-\nabla.

Proof sketch.

We have already seen that the map is an isomorphism of vector spaces, and the Lie brackets are compared in the same way as Lemma 5.21.

It remains to verify that the differentials match up. If GG is a connected graph containing tadpoles (edges connecting a vertex to itself), then both d([G,<])\partial_{d}([G,<]) and c([G,<])\partial_{c}([G,<]) contain terms involving deleting (equivalently, collapsing) those tadpoles, and such terms cancel in the difference (cd)([G,<])(\partial_{c}-\partial_{d})([G,<]). Similarly, if GG contains bridges-to-nowhere (edges connecting a 1-valent vertex to the rest of GG) then deleting such edges agrees with collapsing them modulo K2CK^{2}\subset C, so such terms also cancel in (cd)([G,<])Indec(C)=K/K2(\partial_{c}-\partial_{d})([G,<])\in\mathrm{Indec}(C)=K/K^{2} where K=Ker(C)K=\mathrm{Ker}(C\to\mathbb{Q}). Dualizing the signed sum of collapsing edges which are neither tadpoles nor bridges-to-nowhere precisely matches with the differential δ\delta on 𝖿𝖦𝖢2,conn\mathsf{fGC}^{\circlearrowleft}_{2,\mathrm{conn}}, and dualizing the signed sum of deleting such edges matches with \nabla. ∎

We can then use the results of [KWŽ17] to deduce results about differentials in our spectral sequences.

Proposition 5.24.

The homology group

H(Indec(C))H_{*}(\mathrm{Indec}(C))

with respect to the boundary map induced by =cd\partial=\partial_{c}-\partial_{d} is 1-dimensional in homological degree 11, generated by the class [T][T] of a tadpole graph, and is zero in all other degrees.

Proof.

We consider the composition

(𝖿𝖦𝖢2,conn,δ+)(𝖿𝖦𝖢2,connT,δ+)(𝖿𝖦𝖢2,conn,δ+)Indec(C),(\mathsf{fGC}_{2,\mathrm{conn}},\delta+\nabla)\hookrightarrow(\mathsf{fGC}_{2,\mathrm{conn}}\oplus\mathbb{Q}T^{\vee},\delta+\nabla)\stackrel{{\scriptstyle\simeq}}{{\hookrightarrow}}(\mathsf{fGC}^{\circlearrowleft}_{2,\mathrm{conn}},\delta+\nabla)\cong\mathrm{Indec}(C)^{\vee},

where the last isomorphism is by Proposition 5.23. We justify the middle quasi-isomorphism as follows: by [Wil15, Proposition 3.4], the inclusion

(𝖿𝖦𝖢2,connT,δ)(𝖿𝖦𝖢2,conn,δ)(\mathsf{fGC}_{2,\mathrm{conn}}\oplus\mathbb{Q}T^{\vee},\delta)\xrightarrow{\sim}(\mathsf{fGC}^{\circlearrowleft}_{2,\mathrm{conn}},\delta)

is a quasi-isomorphism, and hence, by a spectral sequence argument (filtering by number of vertices), it follows that

(𝖿𝖦𝖢2,connT,δ+)(𝖿𝖦𝖢2,conn,δ+)(\mathsf{fGC}_{2,\mathrm{conn}}\oplus\mathbb{Q}T^{\vee},\delta+\nabla)\xrightarrow{\sim}(\mathsf{fGC}^{\circlearrowleft}_{2,\mathrm{conn}},\delta+\nabla)

is also a quasi-isomorphism. Finally, [KWŽ17, §3.2, Corollary 4] shows that (𝖿𝖦𝖢2,conn,δ+)(\mathsf{fGC}_{2,\mathrm{conn}},\delta+\nabla) is acyclic, and the claim follows. ∎

Corollary 5.25.

Let

h:Es,trGEs,trQh\colon{}^{G}\!E^{r}_{s,t}\to{}^{Q}\!E^{r}_{s,t}

be the map of spectral sequences induced by GH1(G;)G\mapsto H_{1}(G;\mathbb{Z}). For odd g3g\geq 3, the image

h([Wg])Indec(Eg,g1Q)h([W_{g}])\in\mathrm{Indec}({}^{Q}\!E^{1}_{g,g})

is a permanent cycle in the spectral sequence obtained by taking indecomposables of the Quillen spectral sequence.

Proof.

It follows from [KWŽ17, Proposition 5] that the classes [Wg]Indec(Eg,g1G)[W_{g}]\in\mathrm{Indec}({}^{G}\!E^{1}_{g,g}) survive to represent elements on the Eg1E^{g-1}-page, where a non-zero differential takes the class of [Wg][W_{g}] to the loop class L2g1E1,2g2g1GL_{2g-1}\in{}^{G}\!E^{g-1}_{1,2g-2}. Therefore the image of [Wg][W_{g}] in Indec(Eg,g1Q)\mathrm{Indec}({}^{Q}\!E^{1}_{g,g}) also survives to represent an element of Indec(Eg,gg1Q)\mathrm{Indec}({}^{Q}\!E^{g-1}_{g,g}) which the differential dg1d^{g-1} takes to an element of bidegree (1,2g2)(1,2g-2). But the Quillen spectral sequence is zero in this bidegree, so there are no possible non-zero differentials out of h([Wg])Indec(Eg,grQ)h([W_{g}])\in\mathrm{Indec}({}^{Q}\!E^{r}_{g,g}) for any rr. ∎

As we will explain in the next section, the integration pairing with ω2g1\omega^{2g-1} can be used to deduce that h([Wg])0Indec(Eg,g1Q)h([W_{g}])\neq 0\in\mathrm{Indec}({}^{Q}\!E^{1}_{g,g}). At present we cannot rule out that this class h([Wg])h([W_{g}]) in the corollary above could be in the image of a non-zero differential on some page of the spectral sequence, although this would be impossible assuming the conjecture of [CFP14, §2] which implies that Es,t1Q=0{}^{Q}\!E^{1}_{s,t}=0 for s>ts>t and (s,t)(1,0)(s,t)\neq(1,0). Subject to that conjecture, we can therefore deduce that h([Wg])h([W_{g}]) survives to represent non-zero elements of Indec(H(BK();))=K1()\mathrm{Indec}(H_{*}(BK(\mathbb{Z});\mathbb{Q}))=K_{*-1}(\mathbb{Z})\otimes\mathbb{Q} in degrees {6,10,14,}\ast\in\{6,10,14,\dots\}.

5.6. Inverting the class of a point

We have explained why the spectral sequence E,rG{}^{G}\!E^{r}_{*,*} comes with a product and coproduct, satisfying a Leibniz rule on each page. Furthermore, it is easy to see that there are unit and counit morphisms E0,01G\mathbb{Q}\to{}^{G}\!E^{1}_{0,0}\to\mathbb{Q} induced by filtered maps {}MGr{}\{\ast\}\to M_{\mathrm{Gr}}\to\{\ast\}, making this a spectral sequence of bialgebras. As it stands, the E1E^{1}-page does not admit an antipode though. In this subsection, which is not strictly necessary for the rest of the paper, we investigate the ring E0,01GE0,0rG{}^{G}\!E^{1}_{0,0}\cong{}^{G}\!E^{r}_{0,0} and its action on E,rG{}^{G}\!E^{r}_{*,*} by multiplication.

Lemma 5.26.

The space F0MGrF_{0}M_{\mathrm{Gr}} consists of two path components, both of which have trivial rational homology. Consequently, E0,01G2{}^{G}\!E^{1}_{0,0}\cong\mathbb{Q}^{2} and E0,t1G=0{}^{G}\!E^{1}_{0,t}=0 for t>0t>0.

Proof.

We may instead calculate the homology of the chain complex F0CCF_{0}C\subset C whose generators [G,ω][G,\omega] have b1(G)=0b_{1}(G)=0. In other words, GG is a forest. H0(F0C)H_{0}(F_{0}C) may be calculated by hand: if we write pnp^{n} for the graph with no edges and n0n\geq 0 many vertices, then these form a basis for F0C0F_{0}C_{0}. There is a unique isomorphism class of 1-edge graphs with nn many vertices and b1=0b_{1}=0 for each n2n\geq 2, and its boundary is pn1pnp^{n-1}-p^{n} This shows that the space F0MGrF_{0}M_{\mathrm{Gr}} is the disjoint union of precisely two path components, one corresponding to the empty graph and one corresponding to all non-empty graphs.

To see that Hj(F0C)=0H_{j}(F_{0}C)=0 for j>0j>0 we consider the homomorphism T:F0CnF0Cn+1T\colon F_{0}C_{n}\to F_{0}C_{n+1} given by

(83) T([G,ω])=G=G1G2[G1eG2],T([G,\omega])=\sum_{G=G_{1}\vee G_{2}}[G_{1}\vee e\vee G_{2}],

where the sum is indexed by all possible ways of writing G=G1G2G=G_{1}\vee G_{2} as the wedge sum of two graphs G1G_{1} and G2G_{2} at basepoints v1V(G1)v_{1}\in V(G_{1}) and v2V(G2)v_{2}\in V(G_{2}), with E(G1)E(G2)E(G_{1})\neq\emptyset\neq E(G_{2}). The graph G1eG2G_{1}\vee e\vee G_{2} is obtained from the disjoint union G1G2G_{1}\sqcup G_{2} by inserting a new edge ee connecting v1v_{1} to v2v_{2}, and E(G1eG2)={e}E(G1)E(G2)={e}E(G)E(G_{1}\vee e\vee G_{2})=\{e\}\sqcup E(G_{1})\sqcup E(G_{2})=\{e\}\sqcup E(G) is ordered as the concatenation {e}(E(G),<)\{e\}\ast(E(G),<). Then TT is a chain homotopy from the identity to the chain map S:F0CF0CS\colon F_{0}C\to F_{0}C defined by

S([G,ω])=(T+Tid)([G,ω]).S([G,\omega])=(\partial\circ T+T\circ\partial-\mathrm{id})([G,\omega]).

If GG is a forest all of whose trees have at most nn edges for n2n\geq 2, then S([G,ω])S([G,\omega]) is a linear combination of forests all of whose trees have at most n1n-1 edges. In other words, SS is a chain homotopy from the identity map to a chain map which strictly decreases the maximal number of edges in trees for any tree with more than 1 edge. We now consider a map of CDGAs

[p,e]F0C\mathbb{Q}[p,e]\to F_{0}C

whose domain is the free graded-commutative algebra on a generator pp of degree 0 and a generator ee of degree 1, with boundary defined by p=0\partial p=0 and e=pp2\partial e=p-p^{2}. The map is defined by sending pp to a one-vertex graph and ee to a graph with two vertices connected by an edge. Iterating the chain homotopy SS shows that any cycle in F0CF_{0}C is homologous to a cycle in the image of this map of CDGAs. Finally, an easy calculation shows that Hj([p,e],)=0H_{j}(\mathbb{Q}[p,e],\partial)=0 for j>0j>0. See also [Wil15, Proposition 3.4] for a related result. ∎

Corollary 5.27.

Let xE0,01G=H0(F0MGr;)x\in{}^{G}\!E^{1}_{0,0}=H_{0}(F_{0}M_{\mathrm{Gr}};\mathbb{Q}) be the path component corresponding to the non-empty graphs. Then E0,01G=E0,0G[x]/(x2x){}^{G}\!E^{1}_{0,0}={}^{G}\!E^{\infty}_{0,0}\cong\mathbb{Q}[x]/(x^{2}-x) as a ring, with coproduct given by Δ(x)=xx\Delta(x)=x\otimes x and counit by x1x\mapsto 1.

As already explained, this bialgebra represents the functor R{xRx2=x}R\mapsto\{x\in R\mid x^{2}=x\}, which is a monoid scheme but not a group scheme, so the bialgebra does not admit an antipode.

Proof.

We have already seen that the ring structure is as stated, with x=[p]x=[p] in the notation of the above proof. The equation Δ(x)=xx\Delta(x)=x\otimes x follows from the fact that the coproduct on the EE^{\infty}-page is induced by the space-level diagonal map. Similarly for the counit. ∎

The ring structure on E0,01E0,0rE^{1}_{0,0}\cong E^{r}_{0,0} implies a splitting Es,trxEs,tr(1x)Es,trE^{r}_{s,t}\cong xE^{r}_{s,t}\oplus(1-x)E^{r}_{s,t} for all bidegrees (s,t)(s,t). The following lemma implies that the second summand is trivial for (s,t)(0,0)(s,t)\neq(0,0) and that multiplication by xx acts as the identity in such bidegrees. Formally inverting xx therefore changes only bidegree (0,0)(0,0), and by exactness of localization results in a spectral sequence, which is now a spectral sequence of connected Hopf algebras.

Lemma 5.28.

Let xE0,01G=E0,0rGx\in{}^{G}\!E^{1}_{0,0}={}^{G}\!E^{r}_{0,0} be as in Corollary 5.27. Then multiplication by x:Es,trGEs,trGx\colon{}^{G}\!E^{r}_{s,t}\to{}^{G}\!E^{r}_{s,t} is the identity homomorphism for (s,t)(0,0)(s,t)\neq(0,0) and all r1r\geq 1.

Proof.

It suffices to prove the statement for r=1r=1 where Es,t1G=Hs+t(Grs(C)){}^{G}\!E^{1}_{s,t}=H_{s+t}(\operatorname{Gr}_{s}(C)), since each isomorphism Er+1GH(ErG,dr){}^{G}\!E^{r+1}\cong H_{*}({}^{G}\!E^{r},d^{r}) is an isomorphism of algebras, indeed bialgebras (Corollary 5.9). The map Es,t1GEs,t1G{}^{G}\!E^{1}_{s,t}\to{}^{G}\!E^{1}_{s,t} which multiplies by xx has 1-dimensional kernel and cokernel for (s,t)=(0,0)(s,t)=(0,0), we wish to show that the kernel and cokernel vanishes for s+t>0s+t>0. To see this we look at the quotient complex Gr(C)/(x)\operatorname{Gr}(C)/(x) by the ideal (x)Gr(C)(x)\subset\operatorname{Gr}(C), whose homology sits in a long exact sequence associated to the short exact sequence

(84) 0Gr(C)xGr(C)Gr(C)/(x)0.0\to\operatorname{Gr}(C)\xrightarrow{\cdot x}\operatorname{Gr}(C)\to\operatorname{Gr}(C)/(x)\to 0.

An additive generator [G,<]Gr(C)[G,<]\in\operatorname{Gr}(C) is in (x)(x) if and only if GG contains a vertex of valence 0, while those basis elements [G,<]Gr(C)[G,<]\in\operatorname{Gr}(C) for which all vertices have positive valence reduce to basis elements for Gr(C)/(x)\operatorname{Gr}(C)/(x). In the latter case, we shall use the same notation [G,<][G,<] for the class modulo (x)(x). There are two evident cycles in Gr(C)/(x)\operatorname{Gr}(C)/(x), namely 1Gr(C)/(x)1\in\operatorname{Gr}(C)/(x) represented by the empty graph, and [I,<][I,<] represented by the “interval” graph with two 1-valent vertices connected by a single edge. Both represent non-zero homology classes, and note that the image of the interval graph under the connecting homomorphism H1(Gr(C)/(x))H0(Gr(C))[x]/(x2x)H_{1}(\operatorname{Gr}(C)/(x))\to H_{0}(\operatorname{Gr}(C))\cong\mathbb{Q}[x]/(x^{2}-x) is 1x01-x\neq 0. To prove the Lemma, it will suffice to show that H(Gr(C)/(x))H_{*}(\operatorname{Gr}(C)/(x)) is 1-dimensional in homological degrees 0 and 1, spanned by these two classes, and is zero-dimensional in all other degrees. Indeed, by inspecting the long exact sequence associated to (84), one sees that multiplication by xx is an isomorphism on Hs+t(Gr(C))H_{s+t}(\operatorname{Gr}(C)) for s+t>0s+t>0, but it must then be the identity since we also know that xE0,01Gx\in{}^{G}\!E^{1}_{0,0} is idempotent.

To show that H(Gr(C)/(x))H_{*}(\operatorname{Gr}(C)/(x)) is spanned by the homology classes of 11 and [I,<][I,<], we follow a similar strategy to the proof of Lemma 5.26, but using a different homotopy. For an additive generator [G,<]Gr(C)/(x)[G,<]\in\operatorname{Gr}(C)/(x), i.e., where GG has no zero-valent vertices, we introduce the notation V¯(G)V(G)\overline{V}(G)\subset V(G) for the set of vertices of valence 2\geq 2 and E¯(G)E(G)\overline{E}(G)\subset E(G) for the set of edges which connect two distinct elements of V¯(G)\overline{V}(G). Define a linear homomorphism

(Gr(C)/(x))n𝑇(Gr(C)/(x))n+1(\operatorname{Gr}(C)/(x))_{n}\xrightarrow{T}(\operatorname{Gr}(C)/(x))_{n+1}

by the formula

T[G,<]=vV¯(G)[IvG,<]T[G,<]=\sum_{v\in\overline{V}(G)}[I\vee_{v}G,<^{\prime}]

where GG is a graph with no vertices of valence 0 and IvGI\vee_{v}G denotes the graph obtained from GG by adding a new vertex v0v_{0} and connecting it to vGv\in G by an edge e0e_{0}, so V(IvG)={v0}V(G)V(I\vee_{v}G)=\{v_{0}\}\sqcup V(G) and E(IvG)={e0}E(G)E(I\vee_{v}G)=\{e_{0}\}\sqcup E(G). The total order <<^{\prime} is obtained from << by declaring e0<ee_{0}<^{\prime}e for eE(G)e\in E(G) and agreeing with << on elements of E(G)E(IvG)E(G)\subset E(I\vee_{v}G). Then one verifies that

(Tc+cT)([G,<])=|V¯(G)|[G,<]+eE¯(G)±[Ie/e(G/e),<](T\partial_{c}+\partial_{c}T)([G,<])=|\overline{V}(G)|[G,<]+\sum_{e\in\overline{E}(G)}\pm[I\vee_{e/e}(G/e),<^{\prime}]

where e/eV(G/e)e/e\in V(G/e) is the vertex arising at the collapsed edge. We shall not need a precise recipe for the sign and the induced ordering <<^{\prime}, but note the inequality

(85) |V¯(Ie/e(G/e))|<|V¯(G)|.|\overline{V}(I\vee_{e/e}(G/e))|<|\overline{V}(G)|.

By a similar argument

(Td+dT)([G,<])=(e,v)±[Iv(Ge),<],(T\partial_{d}+\partial_{d}T)([G,<])=\sum_{(e,v)}\pm[I\vee_{v}(G\setminus e),<^{\prime}],

where the sum runs over pairs consisting of an eE¯(G)e\in\overline{E}(G) which is a bridge in GG, and a 2-valent vertex vV(G)v\in V(G) adjacent to ee. Here Iv(Ge)I\vee_{v}(G\setminus e) denotes the result of removing ee and attaching a new edge at vv whose other endpoint is a new 1-valent vertex. Here we note the inequalities

(86) |V¯(Iv(Ge)|\displaystyle|\overline{V}(I\vee_{v}(G\setminus e)| |V¯(G)|,\displaystyle\leq|\overline{V}(G)|,
|E¯(Iv(Ge))|\displaystyle|\overline{E}(I\vee_{v}(G\setminus e))| <|E¯(G)|.\displaystyle<|\overline{E}(G)|.

Now, for integers a,b0a,b\geq 0, let us write DaGr(C)/(x)D_{a}\subset\operatorname{Gr}(C)/(x) for the rational span of those [G,<][G,<] for which |V¯(G)|a|\overline{V}(G)|\leq a and Da,bDaD_{a,b}\subset D_{a} for the rational span of those [G,<][G,<] for which either |V¯(G)|<a|\overline{V}(G)|<a, or |V¯(G)|=a|\overline{V}(G)|=a and |E¯(G)|b|\overline{E}(G)|\leq b. For a>0a>0 let Sa:Gr(C)/(x)Gr(C)/(x)S_{a}\colon\operatorname{Gr}(C)/(x)\to\operatorname{Gr}(C)/(x) be the chain map

Sa=id1a(T+T).S_{a}=\mathrm{id}-\tfrac{1}{a}(T\partial+\partial T).

The inequalities (85) and (86) imply that

Sa(Da,b)\displaystyle S_{a}(D_{a,b}) Da,b1\displaystyle\subset D_{a,b-1}
Sa(Da,0)\displaystyle S_{a}(D_{a,0}) Da1,\displaystyle\subset D_{a-1},

and therefore, if cDac\in D_{a} is any chain and a>0a>0, there exists some integer N0N\geq 0 such that SaNcDa1S_{a}^{N}c\in D_{a-1}. We can regard the linear map 1aT\frac{1}{a}T as a chain homotopy between the identity and SaS_{a}, and by induction we deduce that any chain cGr(C)/(x)c\in\operatorname{Gr}(C)/(x) is homologous to some cD0c^{\prime}\in D_{0}.

Graphs GG with V¯(G)=\overline{V}(G)=\emptyset are easy to classify: each connected component must be isomorphic to the “interval” II, the graph consisting of two vertices connected by a single edge. If GG has more than two such components, then [G,<]=0[G,<]=0 because Aut(G)\mathrm{Aut}(G) contains an odd permutation. We have therefore shown that the inclusion

.1.[I,<]D0Gr(C)/(x)\mathbb{Q}.1\oplus\mathbb{Q}.[I,<]\cong D_{0}\hookrightarrow\operatorname{Gr}(C)/(x)

induces a surjection on homology. ∎

Corollary 5.29.

The rational homology of MGrM_{\mathrm{Gr}} is

Hn(MGr;){2for n=0for n=10for n>1.H_{n}(M_{\mathrm{Gr}};\mathbb{Q})\cong\begin{cases}\mathbb{Q}^{2}&\text{for $n=0$}\\ \mathbb{Q}&\text{for $n=1$}\\ 0&\text{for $n>1$}.\end{cases}

In other words, MGrM_{\mathrm{Gr}} and hence BKGrBK_{\mathrm{Gr}} are rationally equivalent to the disjoint union of a point, corresponding to G=G=\emptyset, and a circle. It seems interesting to understand these homotopy types integrally.

Proof.

We have seen splittings Es,trG=xEs,tr(1x)Es,tr{}^{G}\!E^{r}_{s,t}=xE^{r}_{s,t}\oplus(1-x)E^{r}_{s,t}, in which the second summand is a 1-dimensional rational vector space for (s,t)=(0,0)(s,t)=(0,0) and vanishes in all other bidegrees. The first summand is

xEs,trx1Es,tr,xE^{r}_{s,t}\cong x^{-1}E^{r}_{s,t},

which form the pages of a spectral sequence of connected Hopf algebras. In Subsection 5.5, we identified Indec(x1Es,tr)Indec(Es,trG)\mathrm{Indec}(x^{-1}E^{r}_{s,t})\cong\mathrm{Indec}({}^{G}\!E^{r}_{s,t}) with the pages of the spectral sequence in [KWŽ17], and for r=r=\infty we in particular get that this is 1-dimensional in bidegree (1,0)(1,0) and vanishes otherwise. Since a connected commutative Hopf algebra is determined by its indecomposables, we deduce that x1E,x^{-1}E^{\infty}_{*,*} is an exterior algebra generated by one class in bidegree (1,0)(1,0), and hence

xEs,t{for (s,t)=(0,0) and (s,t)=(1,0)0otherwise.xE^{\infty}_{s,t}\cong\begin{cases}\mathbb{Q}&\text{for $(s,t)=(0,0)$ and $(s,t)=(1,0)$}\\ 0&\text{otherwise.}\end{cases}

Combining with (1x)E0,0(1-x)E^{\infty}_{0,0} gives the result. ∎

6. Freeness of a Lie algebra generated by ω4k+1\omega^{4k+1} classes

We now prove that the images of the elements ω5,ω9,,ω45\omega^{5},\omega^{9},\ldots,\omega^{45} generate a free Lie subalgebra in Prim(W0Hc(𝒜)).\operatorname{Prim}(W_{0}H^{*}_{c}(\mathcal{A})\otimes\mathbb{R}). We will make use of the map (82) from the primitives in the Quillen spectral sequence (along diagonal bidegrees) to H0(𝖦𝖢2)H^{0}(\mathsf{GC}_{2}) constructed in the previous section, as well as the following results from the literature:

  1. (1)

    Willwacher’s theorem [Wil15, Theorem 1.1] which gives an isomorphism of graded Lie algebras between H0(𝖦𝖢2)H^{0}(\mathsf{GC}_{2}) and the Grothendieck–Teichmüller Lie algebra, which we shall denote by 𝔤𝔯𝔱\mathfrak{grt} (denoted by 𝔤𝔯𝔱1\mathfrak{grt}_{1} in loc. cit.).

  2. (2)

    An injective map [Bro12] from the motivic Lie algebra 𝔤𝔪𝔤𝔯𝔱\mathfrak{g}^{\mathfrak{m}}\to\mathfrak{grt}, where 𝔤𝔪\mathfrak{g}^{\mathfrak{m}} is isomorphic to the free graded Lie algebra on certain generators σ2k+1\sigma_{2k+1} in every degree 2k+132k+1\geq 3. These generators σ2k+1\sigma_{2k+1} are canonical modulo Lie words in generators σ2j+1\sigma_{2j+1} with j<kj<k.

  3. (3)

    The integration pairing between ω2g1E1g,gQ\omega^{2g-1}\in{}^{Q}\!E^{g,g}_{1}\otimes\mathbb{R} (see (32)) and the locally-finite homology class of the wheel [Wg][W_{g}] is non-zero for g>1g>1 odd [BS24]. In other words, the image of ω2g1\omega^{2g-1} in H0(𝖦𝖢2)H^{0}(\mathsf{GC}_{2})\otimes\mathbb{R}, under the map (82), pairs non-trivially with [Wg]H0(𝖦𝖢2)[W_{g}]\in H_{0}(\mathsf{GC}_{2}^{\vee}).

This last point (3) relies on the discussion of subsection 2.4, in the following way. The differential form ω2g1\omega^{2g-1} most naturally gives a class in the E1E_{1}-page of the cohomological tropical spectral sequence, while the wheel graph WgW_{g} gives a class in the homological Quillen spectral sequence. In order to make sense of the pairing, we use the explicit comparison discussed in §2.4, which we recall involved a zig-zag

Grg(BK())|NT(g){}|(Pg/GLg()){}.\operatorname{Gr}_{g}(BK(\mathbb{Z}))\xleftarrow{\simeq}|N_{\bullet}T_{\bullet}(\mathbb{Q}^{g})\cup\{\infty\}|\xrightarrow{\simeq_{\mathbb{Q}}}(P_{g}/\mathrm{GL}_{g}(\mathbb{Z}))\cup\{\infty\}.

We use this to make sense of pairing a reduced homology class in Grg(BK())\operatorname{Gr}_{g}(BK(\mathbb{Z})) with a reduced cohomology class in (Pg/GLg()){}(P_{g}/\mathrm{GL}_{g}(\mathbb{Z}))\cup\{\infty\}, or equivalently a compactly supported cohomology class in Pg/GLg()P_{g}/\mathrm{GL}_{g}(\mathbb{Z}). To implement the comparison, we will first factor the second arrow through a space |T¯(g){}||\overline{T}_{\bullet}(\mathbb{Q}^{g})\cup\{\infty\}| which we define now. Let T¯p(g)\overline{T}_{p}(\mathbb{Q}^{g}) be the quotient of the set N0Tp(g)N_{0}T_{p}(\mathbb{Q}^{g}) by the equivalence relation that (ApA0,<)(ApA0,<)(A_{p}\subset\dots\subset A_{0},<)\sim(A^{\prime}_{p}\subset\dots\subset A^{\prime}_{0},<^{\prime}) provided the resulting maps

ΔpΔpPg,\Delta^{p}\setminus\partial\Delta^{p}\to P_{g},

given by the formula (25) are in the same orbit for the action of GLg()\mathrm{GL}_{g}(\mathbb{Z})-action on the set of such maps. We point out that the formula (25) for the map associated to a=(ApA0,<)a=(A_{p}\subset\dots\subset A_{0},<) does not involve the total order << at all, and that the elements ψA0\psi\in A_{0} enter the formula only through their squares v(ψ(v))2v\mapsto(\psi(v))^{2}. Therefore the element of T¯p(g)\overline{T}_{p}(\mathbb{Q}^{g}) represented by aa is invariant under any change of <<, under replacing some ψ\psi by ψ-\psi, and by definition also under precomposing all ψ\psi’s with the same XGLg()X\in\mathrm{GL}_{g}(\mathbb{Z}).

The definition of the equivalence relation defining T¯p(g)\overline{T}_{p}(\mathbb{Q}^{g}) ensures that the map from Proposition 2.20 factors as

Grg(BK())|NT(g){}||T¯(g){}|(Pg/GLg()){}.\operatorname{Gr}_{g}(BK(\mathbb{Z}))\xleftarrow{\simeq}|N_{\bullet}T_{\bullet}(\mathbb{Q}^{g})\cup\{\infty\}|\to|\overline{T}_{\bullet}(\mathbb{Q}^{g})\cup\{\infty\}|\to(P_{g}/\mathrm{GL}_{g}(\mathbb{Z}))\cup\{\infty\}.

We may also define a map of simplicial sets

Grg()𝑓T¯(g){}\operatorname{Gr}_{g}(\mathcal{F}_{\bullet})\xrightarrow{f}\overline{T}_{\bullet}(\mathbb{Q}^{g})\cup\{\infty\}

by sending the isomorphism class of a flag x=(G0,0G0,p)Fgpx=(G_{0,0}\hookrightarrow\dots\hookrightarrow G_{0,p})\in F_{g}\mathcal{F}_{p} to the element of Tp(g){}T_{p}(\mathbb{Q}^{g})\cup\{\infty\} given as f(x)=f(x)=\infty if b1(G0,p)<gb_{1}(G_{0,p})<g and otherwise by first choosing an isomorphism H1(G0,p;)gH_{1}(G_{0,p};\mathbb{Z})\cong\mathbb{Z}^{g} and then letting

f(x)=[(ApA0,<)]T¯p(g),f(x)=[(A_{p}\subset\dots\subset A_{0},<)]\in\overline{T}_{p}(\mathbb{Q}^{g}),

where A0H1(G0,p;)(g)A_{0}\subset H^{1}(G_{0,p};\mathbb{Q})\cong(\mathbb{Q}^{g})^{\vee} is the subset defined by first choosing a section of H(G0,p)E(G0,p)H(G_{0,p})\twoheadrightarrow E(G_{0,p}) and then letting A0A_{0} consist of all non-zero elements in the image of the resulting composition

E(G0,p)H(G0,p)C1(G;)H1(G;)g.E(G_{0,p})\hookrightarrow H(G_{0,p})\hookrightarrow C^{1}(G;\mathbb{Q})\twoheadrightarrow H^{1}(G;\mathbb{Q})\cong\mathbb{Q}^{g}.

We remark that an edge is sent to zero precisely when it is a bridge. Let AiA0A_{i}\subset A_{0} consist of the non-zero vectors in the image of the subset E(G0,pi)E(G0,p)E(G_{0,p-i})\subset E(G_{0,p}) for i=0,,pi=0,\dots,p and choose an arbitrary total ordering << of A0A_{0} to define the element f(x)f(x). (The resulting element f(x)f(x) does not depend on the choice of <<, nor on the choice of representative half-edges of each edge, by definition of the equivalence relation defining T¯(g)\overline{T}_{\bullet}(\mathbb{Q}^{g}).) These maps fit into a homotopy commutative diagram

Grg(BKGr){{\operatorname{Gr}_{g}(BK_{\mathrm{Gr}})}}Grg(BK()){{\operatorname{Gr}_{g}(BK(\mathbb{Z}))}}|NT(g){}|{{|N_{\bullet}T_{\bullet}(\mathbb{Q}^{g})\cup\{\infty\}|}}Grg(MGr){{\operatorname{Gr}_{g}(M_{\mathrm{Gr}})}}|T¯(g){}|{{|\overline{T}_{\bullet}(\mathbb{Q}^{g})\cup\{\infty\}|}}(Pg/GLg()){}.{{(P_{g}/\mathrm{GL}_{g}(\mathbb{Z}))\cup\{\infty\}.}}H1\scriptstyle{H_{1}}\scriptstyle{\simeq_{\mathbb{Q}}}\scriptstyle{\simeq}\scriptstyle{\simeq_{\mathbb{Q}}}f\scriptstyle{f}

Now, a graph GG with b1(G)=gb_{1}(G)=g defines a map

(Δ1)E(G)Grg(¯)(\Delta^{1}_{\bullet})^{E(G)}\to\operatorname{Gr}_{g}(\overline{\mathcal{F}}_{\bullet})

whose geometric realization is one of the cells in the rational cell structure leading to the cellular chain complex Grg(C)\operatorname{Gr}_{g}(C) calculating the reduced homology of Grg(MGr)\operatorname{Gr}_{g}(M_{\mathrm{Gr}}) which is the column Eg,1G{}^{G}\!E^{1}_{g,*} in the graph spectral sequence. Composing with the bottom row of the diagram above, we obtain a map

(Δ1)E(G)(Pg/GLg()){}.(\Delta^{1})^{E(G)}\to(P_{g}/\mathrm{GL}_{g}(\mathbb{Z}))\cup\{\infty\}.

Reparametrizing using the inverse 1:(0,)Δ1Δ1\ell^{-1}\colon(0,\infty)\to\Delta^{1}\setminus\partial\Delta^{1} of the diffeomorphism (26), the composition

(0,)E(G)(1)E(G)(Δ1)E(G)(Pg/GLg()){}(0,\infty)^{E(G)}\xrightarrow{(\ell^{-1})^{E(G)}}(\Delta^{1})^{E(G)}\to(P_{g}/\mathrm{GL}_{g}(\mathbb{Z}))\cup\{\infty\}

is independent of choice of \ell. Tracing through definitions, it may be identified with the graph Laplacian (0,)E(G)Pg(0,\infty)^{E(G)}\to P_{g} from [BS24], which depends on a choice of basis of H1(G;)H_{1}(G;\mathbb{Z}), composed with the quotient map PgPg/GLg()P_{g}\to P_{g}/\mathrm{GL}_{g}(\mathbb{Z}).

For g=2k+1g=2k+1, integrating the invariant differential form ω4k+1\omega^{4k+1} along this map for each GG with |E(G)|=4k+2|E(G)|=4k+2 defines a element of degree 4k+24k+2 in the linear dual cochain complex

ω4k+1Grg(C)=Hom(Grg(C),).\omega^{4k+1}\in\operatorname{Gr}_{g}(C)^{\vee}=\operatorname{Hom}(\operatorname{Gr}_{g}(C),\mathbb{Q}).

This element is a cocycle and pairs with the cycle [W2k+1]Grg(C)[W_{2k+1}]\in\operatorname{Gr}_{g}(C) by evaluating the integral from [BS24]. By the discussion above, this is the desired pairing.

6.1. Proof of Theorem 1.3

Lemma 6.1.

For all g>1g>1 odd, the wheel classes [Wg]Eg,g1G[W_{g}]\in{}^{G}\!E^{1}_{g,g} are primitive, and hence annihilated by the Lie cobracket. Consequently, the wheel classes in H0(𝖦𝖢2)H_{0}(\mathsf{GC}_{2}^{\vee}) define linear maps

[Wg]:𝔤𝔯𝔱[W_{g}]\colon\mathfrak{grt}\to\mathbb{Q}

which vanish on commutators [𝔤𝔯𝔱,𝔤𝔯𝔱][\mathfrak{grt},\mathfrak{grt}], and hence descend to maps 𝔤𝔯𝔱ab\mathfrak{grt}^{\mathrm{ab}}\to\mathbb{Q}, which are nonzero by (3).

Proof.

The coproduct on E1G{}^{G}\!E^{1} is induced by (74). Since every proper nonempty subgraph of a wheel WgW_{g} has a vertex of degree <3<3 (or, if one prefers, since every non-trivial contraction of a number of edges in a wheel graph produces a doubled edge), it follows that any term γWg/γ\gamma\otimes W_{g}/\gamma in the reduced coproduct is zero in CCC\otimes C. It follows that Δ[Wg]=1[Wg]+[Wg]1\Delta[W_{g}]=1\otimes[W_{g}]+[W_{g}]\otimes 1. As a result, the Lie cobracket on H(L)H_{*}(L^{\vee}), which is obtained by antisymmetrizing Δ\Delta, vanishes on [Wg][W_{g}]. ∎

The fact that the wheel class [W2k+1]H0(𝖦𝖢2)[W_{2k+1}]\in H_{0}(\mathsf{GC}_{2}^{\vee}) is non-zero in Hom(𝔤𝔯𝔱ab,)\mathrm{Hom}(\mathfrak{grt}^{\mathrm{ab}},\mathbb{Q}) may also be deduced from a theorem of Rossi and Willwacher [RW14], asserting that it pairs non-trivially with the image of σ2k+1\sigma_{2k+1} in 𝔤𝔯𝔱H0(𝖦𝖢2)\mathfrak{grt}\cong H^{0}(\mathsf{GC}_{2}) for every k>1k>1.

Proposition 6.2.

For g>1g>1 odd, the image of the canonical class [ω2g1]E1g,gQ[\omega^{2g-1}]\in{}^{Q}\!E_{1}^{g,g}\otimes_{\mathbb{Q}}\mathbb{R} under the map (82) has non-zero image in the abelianization 𝔤𝔯𝔱ab\mathfrak{grt}^{\mathrm{ab}}\otimes_{\mathbb{Q}}\mathbb{R}.

Proof.

Let g>1g>1 and suppose on the contrary that [ω2g1][𝔤𝔯𝔱,𝔤𝔯𝔱][\omega^{2g-1}]\in[\mathfrak{grt},\mathfrak{grt}]\otimes_{\mathbb{Q}}\mathbb{R} lies in the subspace of commutators of H0(𝖦𝖢2)H^{0}(\mathsf{GC}_{2})\otimes_{\mathbb{Q}}\mathbb{R}. Then by the previous lemma, the pairing [ω2g1],[Wg]\langle[\omega^{2g-1}],[W_{g}]\rangle must vanish, contradicting item (3). ∎

We immediately deduce a number of consequences. The weight filtration on 𝔤𝔪\mathfrak{g}^{\mathfrak{m}} and 𝔤𝔯𝔱\mathfrak{grt} is induced by a grading which we denote by WW. For the former, this is defined to be (minus) half the Hodge-theoretic weight. For the latter, 𝔤𝔯𝔱\mathfrak{grt} is by definition embedded as a vector space in the free graded Lie algebra on two generators 𝕃(X,Y)\mathbb{L}(X,Y) where X,YX,Y are assigned weight 11. This weight has the property that any generator σ2k+1𝔤𝔪\sigma_{2k+1}\in\mathfrak{g}^{\mathfrak{m}} has weight 2k+12k+1, and so that the weight coincides with the grading by the genus (or equivalently by one half of the cohomological degree) on the diagonal gE1g,gG\bigoplus_{g}{}^{G}\!E^{g,g}_{1}.

Corollary 6.3.

Let N>0N>0 odd such that GrkW𝔤𝔪=GrkW𝔤𝔯𝔱\operatorname{Gr}^{W}_{k}\mathfrak{g}^{\mathfrak{m}}=\operatorname{Gr}^{W}_{k}\mathfrak{grt} for kNk\leq N. Then the images of

ω5,,ω2N1\omega^{5},\ldots,\omega^{2N-1}

under the map Prim(gE1g,gQ)𝔤𝔯𝔱\mathrm{Prim}(\bigoplus_{g}{}^{Q}\!E_{1}^{g,g}\otimes_{\mathbb{Q}}\mathbb{R})\to\mathfrak{grt}\otimes_{\mathbb{Q}}\mathbb{R} from (82) generate a free Lie algebra in 𝔤𝔯𝔱\mathfrak{grt}\otimes_{\mathbb{Q}}\mathbb{R}. Consequently

T(k=1N12ω4k+1[1])g0E1g,gQT\left(\bigoplus_{k=1}^{\frac{N-1}{2}}\omega^{4k+1}[-1]\mathbb{Q}\right)\to\bigoplus_{g\geq 0}{}^{Q}\!E_{1}^{g,g}\otimes_{\mathbb{Q}}\mathbb{R}

is injective.

Proof.

We have established that for all g>1g>1 odd, the image of [ω2g1][\omega^{2g-1}] in 𝔤𝔯𝔱\mathfrak{grt}\otimes\mathbb{R} is non-trivial in 𝔤𝔯𝔱ab\mathfrak{grt}^{\mathrm{ab}}\otimes\mathbb{R} and has graded weight gg. If gNg\leq N then the image of [ω2g1][\omega^{2g-1}] lies in WN𝔤𝔯𝔱W_{N}\mathfrak{grt}\,\otimes\mathbb{R} and since WN𝔤𝔯𝔱=WN𝔤𝔪W_{N}\mathfrak{grt}=W_{N}\mathfrak{g}^{\mathfrak{m}}, it necessarily lies in the graded Lie subalgebra 𝔤𝔪𝔤𝔯𝔱\mathfrak{g}^{\mathfrak{m}}\otimes\mathbb{R}\subset\mathfrak{grt}\otimes\mathbb{R}, and must have non-zero image in (𝔤𝔪)ab(\mathfrak{g}^{\mathfrak{m}})^{\mathrm{ab}}\otimes\mathbb{R}. By [Bro12], 𝔤𝔪\mathfrak{g}^{\mathfrak{m}} is a free Lie algebra, with generators given by any homogeneous choice of representatives for (𝔤𝔪)ab=H1(𝔤𝔪;)(\mathfrak{g}^{\mathfrak{m}})^{\mathrm{ab}}\otimes\mathbb{R}=H_{1}(\mathfrak{g}^{\mathfrak{m}};\mathbb{R}) in 𝔤𝔪\mathfrak{g}^{\mathfrak{m}}. It follows that ω5,,ω2N1\omega^{5},\ldots,\omega^{2N-1} generate a free graded Lie algebra 𝔤\mathfrak{g}^{\prime} inside 𝔤𝔯𝔱\mathfrak{grt}\otimes\mathbb{R}, and hence inside Prim(E1Q)\mathrm{Prim}({}^{Q}\!E_{1}). The Milnor–Moore theorem then implies that the universal enveloping algebra 𝒰𝔤\mathcal{U}\mathfrak{g}^{\prime} embeds into E1g,gQ\bigoplus{}^{Q}\!E_{1}^{g,g}\otimes_{\mathbb{Q}}\mathbb{R}. The last statement follows from the fact that the universal enveloping algebra of a free graded Lie algebra is isomorphic to the free tensor algebra on its generators. ∎

Establishing Corollary 6.3 for a given NN reduces to a finite, but possibly very large, computation. In practice, it is equivalent to show the equality of finite-dimensional vector spaces WN𝒪(𝔤𝔪)=WN𝒪(𝔤𝔯𝔱)W_{N}\mathcal{O}(\mathfrak{g}^{\mathfrak{m}})=W_{N}\mathcal{O}(\mathfrak{grt}) (where the affine rings 𝒪(𝔤)\mathcal{O}(\mathfrak{g}) are the graded duals of the enveloping algebras 𝒰𝔤\mathcal{U}\mathfrak{g}, for 𝔤=𝔤𝔪,𝔤𝔯𝔱\mathfrak{g}=\mathfrak{g}^{\mathfrak{m}},\mathfrak{grt}) which can be verified by interpreting elements of their affine rings as formal symbols (corresponding to multiple zeta values) modulo relations and checking that their dimensions agree. Indeed, extensive computer calculations of the latter imply that the assumption of the corollary holds for N=23N=23 (see, for example, [BBV10, p. 2], where it was asserted that the dimension was computed up to weight 2424).

One could possibly push this further using more recent results and techniques which exploit the Lie algebra structure. In any case, this corollary may be combined with previous results to produce a huge amount of cohomology: we deduce that the symmetric algebra on Ωc[1]\Omega^{*}_{c}[-1] and the non-trivial commutators in ω5,,ω45\omega^{5},\ldots,\omega^{45} embeds into the E1E_{1}-page of the cohomological Quillen spectral sequence.

6.2. Depth filtration

The de Rham fundamental Lie algebra of the projective line minus 3 points ([Del89], see also [Bro21a] and the references therein) is canonically isomorphic to the free graded Lie algebra 𝕃(X,Y)\mathbb{L}(X,Y) on two generators X,YX,Y (corresponding to generators of HdR1(1\{0,1,};)XYH^{1}_{\mathrm{dR}}(\mathbb{P}^{1}\backslash\{0,1,\infty\};\mathbb{Q})\cong\mathbb{Q}X\oplus\mathbb{Q}Y). The depth filtration on 𝕃(X,Y)\mathbb{L}(X,Y) is the decreasing filtration such that elements of depth rr are linear combinations of Lie brackets involving at least rr occurrences of the letter YY. It induces a decreasing depth filtration DD on both the Grothendieck–Teichmüller and motivic Lie algebras. It is known that

GrD1𝔤𝔪=GrD1𝔤𝔯𝔱,\operatorname{Gr}^{1}_{D}\mathfrak{g}^{\mathfrak{m}}=\operatorname{Gr}^{1}_{D}\mathfrak{grt}\ ,

since both sides of this equation are isomorphic to the graded \mathbb{Q}-vector space generated by ad(X)2n(Y)\mathrm{ad}(X)^{2n}(Y) for n1n\geq 1, and furthermore D1𝔤𝔯𝔱=𝔤𝔯𝔱D^{1}\mathfrak{grt}=\mathfrak{grt}, and [𝔤𝔯𝔱,𝔤𝔯𝔱]D2𝔤𝔯𝔱[\mathfrak{grt},\mathfrak{grt}]\subset D^{2}\mathfrak{grt}. It follows that there is a natural surjective map 𝔤𝔯𝔱abGrD1𝔤𝔯𝔱\mathfrak{grt}^{\mathrm{ab}}\rightarrow\operatorname{Gr}^{1}_{D}\mathfrak{grt}. A question of Drinfeld’s, which asks if 𝔤𝔪𝔤𝔯𝔱\mathfrak{g}^{\mathfrak{m}}\rightarrow\mathfrak{grt} is surjective, would imply that 𝔤𝔯𝔱abGrD1𝔤𝔯𝔱\mathfrak{grt}^{\mathrm{ab}}\rightarrow\operatorname{Gr}^{1}_{D}\mathfrak{grt} is an isomorphism (it is known that GrD1𝔤𝔪=(𝔤𝔪)ab\operatorname{Gr}^{1}_{D}\mathfrak{g}^{\mathfrak{m}}=(\mathfrak{g}^{\mathfrak{m}})^{\mathrm{ab}} is an isomorphism). We have the following stronger version of Proposition 6.2.

Proposition 6.4.

For g>1g>1 odd, the image of the forms [ω2g1][\omega^{2g-1}] are non-zero in GrD1𝔤𝔯𝔱\operatorname{Gr}^{1}_{D}\mathfrak{grt}\otimes_{\mathbb{Q}}\mathbb{R}.

Proof.

The proof of [Wil15, Proposition 9.1] shows that the isomorphism ϕ:H0(𝖦𝖢2)𝔤𝔯𝔱\phi\colon H^{0}(\mathsf{GC}_{2})\rightarrow\mathfrak{grt} constructed in loc. cit. lifts to a map from 𝖦𝖢2\mathsf{GC}_{2} to the free Lie algebra on two generators 𝕃(X,Y)\mathbb{L}(X,Y), and furthermore has the property that the only graph whose image involves the Lie word ad(X)2n(Y)\mathrm{ad}(X)^{2n}(Y) is the wheel W2n+1W_{2n+1}. It follows that the image ϕ(G)\phi(G) for all connected graphs GG not isomorphic to a wheel lies in D2𝕃(X,Y)D^{2}\mathbb{L}(X,Y), since the depth filtration is induced by the degree in YY, and hence GrD1𝕃(X,Y)\operatorname{Gr}^{1}_{D}\mathbb{L}(X,Y) is generated in weight 2n+12n+1 by precisely ad(X)2n(Y)\mathrm{ad}(X)^{2n}(Y). Denote the increasing filtration on H0(𝖦𝖢2)H_{0}(\mathsf{GC}_{2}^{\vee}) dual to the depth filtration on 𝔤𝔯𝔱\mathfrak{grt} by DD_{\bullet}. By the above, it has the property that D1H0(𝖦𝖢2)D_{1}H_{0}(\mathsf{GC}_{2}^{\vee}) is spanned by the wheel classes [W2n+1][W_{2n+1}] for n>1n>1. Since [ω2g1][\omega^{2g-1}] pairs non-trivially with the wheel [Wg][W_{g}], for g>1g>1 odd, it follows that [ω2g1]𝔤𝔯𝔱[\omega^{2g-1}]\in\mathfrak{grt}\otimes_{\mathbb{Q}}\mathbb{R} is not contained in D2D^{2}. ∎

Unfortunately, the bigraded Lie algebra generated by GrD1𝔤𝔯𝔱=GrD1𝔤𝔪\operatorname{Gr}^{1}_{D}\mathfrak{grt}=\operatorname{Gr}^{1}_{D}\mathfrak{g}^{\mathfrak{m}} is not free, and has quadratic relations coming from cusp forms (Ihara–Takao). It is nevertheless very large.

Corollary 6.5.

The image of the Lie algebra generated by {ω5,ω9,}Prim(W0Hc(𝒜;))\{\omega^{5},\omega^{9},\ldots\}\subset\operatorname{Prim}(W_{0}H^{*}_{c}(\mathcal{A};\mathbb{R})) surjects onto the bigraded Lie subalgebra of GrD𝔤𝔪\operatorname{Gr}^{*}_{D}\mathfrak{g}^{\mathfrak{m}} generated by GrD1𝔤𝔪\operatorname{Gr}^{1}_{D}\mathfrak{g}^{\mathfrak{m}} which has one generator in every odd degree >1>1, namely the images of the σ2k+1\sigma_{2k+1} modulo D2.D^{2}.

This result and, for example, the exact sequence [Bro21a, (7.8)] (or more precisely, the dimensions of the graded pieces of the weight-graded vector space denoted by 𝔻2\mathbb{D}_{2} in that paper, which is isomorphic to GrD2𝔤𝔪\operatorname{Gr}^{2}_{D}\mathfrak{g}^{\mathfrak{m}}, which can be traced back to [Zag93], equations (11), (12) and following discussion) implies the following lower bound for the subspace of Prim(W0Hc(𝒜;))\operatorname{Prim}(W_{0}H^{*}_{c}(\mathcal{A};\mathbb{R})) generated by commutators of generators ω4k+1\omega^{4k+1}:

(87) dim(k+=N[ω4k+1,ω4+1])N3.\dim_{\mathbb{R}}\left(\sum_{k+\ell=N}[\omega^{4k+1},\omega^{4\ell+1}]\mathbb{R}\right)\geq\left\lfloor\frac{N}{3}\right\rfloor\ .

We expect that the brackets [ω4k+1,ω4+1][\omega^{4k+1},\omega^{4\ell+1}] for k<k<\ell are linearly independent if the map in Question 1.17 is injective. That expectation, if true, would imply a lower bound for the dimension on the left hand side of (87) of order N/2N/2, while the above argument by depth filtration gives us an unconditional lower bound of order N/3N/3. This lower bound complements the statement of Corollary 6.3. That corollary implies that the ω4k+1\omega^{4k+1} for small kk generate a free Lie algebra, i.e., arbitrary long Lie brackets in ω4k+1\omega^{4k+1} for small kk are independent (modulo antisymmetry and Jacobi relations), but leaves open the possibility that the ω4k+1\omega^{4k+1} could in principle commute for large kk. The lower bound (87) rules out this possibility and proves an orthogonal statement, namely, the independence of many Lie brackets of length two for generators ω4k+1\omega^{4k+1} for arbitrarily large kk. Both results, namely Corollary 6.3 and (87), point to the highly non-commutative nature of the product in E1,Q{}^{Q}\!E^{*,*}_{1} and contribute to the body of evidence in favour of Conjecture 4.6 to be discussed below.

7. Further results and conjectures

7.1. Symmetric products of canonical forms and an announcement of Ronnie Lee

An immediate corollary of Theorem 4.5 is:

Corollary 7.1.

The map (57) induces an embedding of bigraded coalgebras

Sym(Ωc[1])W0Hc(𝒜;)\operatorname{Sym}\left(\Omega^{*}_{c}[-1]\right)\hookrightarrow W_{0}H_{c}^{*}(\mathcal{A};\mathbb{R})

where Sym(Ωc[1])T(Ωc[1])\operatorname{Sym}\left(\Omega^{*}_{c}[-1]\right)\subset T(\Omega^{*}_{c}[-1]) is the vector subspace of (graded-commutative) symmetrized products. By applying the isomorphism W0Hc(𝒜;)[x]/x2E1,QW_{0}H_{c}^{*}(\mathcal{A};\mathbb{Q})\otimes\mathbb{Q}[x]/x^{2}\cong{}^{Q}\!E^{*,*}_{1}, with xx in bidegree (1,0)(1,0), we deduce an embedding of bigraded coalgebras:

(88) Sym(Ωc[1])[x]/x2g,dHcdg(Pg/GLg();).\operatorname{Sym}\left(\Omega^{*}_{c}[-1]\right)\otimes\mathbb{Q}[x]/x^{2}\hookrightarrow\bigoplus_{g,d}H^{d-g}_{c}(P_{g}/\mathrm{GL}_{g}(\mathbb{Z});\mathbb{R})\ .
Proof.

Since W0Hc(𝒜;)W_{0}H_{c}^{*}(\mathcal{A};\mathbb{Q}) is connected, cocommutative, and finite dimensional in each degree, the Milnor–Moore theorem ([MM65, Theorem 5.18]) gives an isomorphism of Hopf algebras

𝒰(Prim(W0Hc(𝒜;)))W0Hc(𝒜;)\mathcal{U}(\mathrm{Prim}(W_{0}H_{c}^{*}(\mathcal{A};\mathbb{Q})))\overset{\sim}{\rightarrow}W_{0}H_{c}^{*}(\mathcal{A};\mathbb{Q})

of bigraded Hopf algebras. The Poincaré–Birkhoff–Witt theorem implies that the natural map Sym(Prim(W0Hc(𝒜;)))Gr𝒰(Prim(W0Hc(𝒜;)))\mathrm{Sym}(\mathrm{Prim}(W_{0}H_{c}^{*}(\mathcal{A};\mathbb{Q})))\rightarrow\operatorname{Gr}\mathcal{U}(\mathrm{Prim}(W_{0}H_{c}^{*}(\mathcal{A};\mathbb{Q}))) is an isomorphism of algebras, where Gr\operatorname{Gr} is the grading associated to the filtration induced by length. In particular, the map given by symmetrisation of products Sym(Prim(W0Hc(𝒜;)))𝒰(Prim(W0Hc(𝒜;)))\mathrm{Sym}(\mathrm{Prim}(W_{0}H_{c}^{*}(\mathcal{A};\mathbb{Q})))\rightarrow\mathcal{U}(\mathrm{Prim}(W_{0}H_{c}^{*}(\mathcal{A};\mathbb{Q}))) is an isomorphism of vector spaces (and in fact of coalgebras if given the appropriate prefactor 1/n!1/n! in length nn; see [Qui69, Theorem B2.3]. Note that it can sometimes be convenient to choose a different splitting of the length filtration, such as in table 1, in which case one only obtains an isomorphism of vector spaces).

Thus we deduce an isomorphism of bigraded coalgebras:

Sym(Prim(W0Hc(𝒜;)))W0Hc(𝒜;)\mathrm{Sym}(\mathrm{Prim}(W_{0}H_{c}^{*}(\mathcal{A};\mathbb{R})))\overset{\sim}{\longrightarrow}W_{0}H_{c}^{*}(\mathcal{A};\mathbb{R})

By Theorem 1.2, Ωc[1]\Omega_{c}^{*}[-1] embeds into Prim(W0Hc(𝒜;))\mathrm{Prim}(W_{0}H_{c}^{*}(\mathcal{A};\mathbb{R})).

The second statement follows from the description of the E1E_{1}-page of the cohomological Quillen spectral sequence in terms of compactly supported cohomology

E1s,tQHt(GLs();Sts)Hcs+t(Ps/GLs();),{}^{Q}\!E_{1}^{s,t}\cong H_{t}(\mathrm{GL}_{s}(\mathbb{Z});\operatorname{St}_{s}\otimes\mathbb{Q})^{\vee}\cong H^{s+t}_{c}(P_{s}/\mathrm{GL}_{s}(\mathbb{Z});\mathbb{Q}),

see (6) and (9). ∎

A version of the following injective map, which is implied by (88),

Sym(Ωc[1])gHc(Pg/GLg();)\operatorname{Sym}\left(\Omega^{*}_{c}[-1]\right)\hookrightarrow\bigoplus_{g}H^{*}_{c}(P_{g}/\mathrm{GL}_{g}(\mathbb{Z});\mathbb{R})

was announced by Ronnie Lee in 1978 [Lee78], but no proof has ever appeared in print. The map (88) gives rise to infinitely many new classes in the cohomology of the groups SLg()\mathrm{SL}_{g}(\mathbb{Z}).

Example 7.2.

By Corollary 7.1, we have the following nonzero classes in Hc(P6/GL6();)H_{c}^{*}(P_{6}/\mathrm{GL}_{6}(\mathbb{Z});\mathbb{R}):

[ω9].ϵ,[ω5].[ω5]=[ω5|ω5],[ω5ω9].ϵ,[\omega^{9}].\epsilon\ ,\qquad[\omega^{5}].[\omega^{5}]=[\omega^{5}|\omega^{5}]\ ,\qquad[\omega^{5}\wedge\omega^{9}].\epsilon\ ,

in degrees 11,12,1611,12,16 respectively. In Hc(P7/GL7();)H_{c}^{*}(P_{7}/\mathrm{GL}_{7}(\mathbb{Z});\mathbb{R}) we obtain:

[ω5|ω5].ϵ,[ω13],[ω5ω13],[ω9ω13],[ω5ω9ω13],[\omega^{5}|\omega^{5}].\epsilon\ ,\quad[\omega^{13}]\ ,\quad[\omega^{5}\wedge\omega^{13}]\ ,\quad[\omega^{9}\wedge\omega^{13}]\ ,\quad[\omega^{5}\wedge\omega^{9}\wedge\omega^{13}]\ ,

in degrees 13,14,19,23,2813,14,19,23,28. Here, a dot denotes the symmetrized product: a.b=12(a×b+(1)degadegbb×a)a.b=\frac{1}{2}(a\times b+(-1)^{\deg a\cdot\deg b}b\times a) where ×\times is the multiplication in E1Q.{}^{Q}\!E_{1}. It follows from the computations of [EVGS13] that these are the only non-vanishing classes in the range for which the compactly supported cohomology of Pg/GLg()P_{g}/\mathrm{GL}_{g}(\mathbb{Z}) has been computed in its entirety.

Remark 7.3.

A much stronger version of the Corollary 7.1 holds. By a similar application of the Milnor–Moore theorem, the symmetric algebra generated by:

  1. (1)

    Independent elements in the Lie algebra generated by {ω5,ω9,}\{\omega^{5},\omega^{9},\ldots\} inside PrimE1Q\mathrm{Prim}{}^{Q}\!E_{1}\otimes\mathbb{R}

  2. (2)

    The homogeneous elements in Ωc[1]\Omega_{c}^{*}[-1] of the form ω4i1+1ω4ik+1\omega^{4i_{1}+1}\wedge\ldots\wedge\omega^{4i_{k}+1} for k>1k>1

  3. (3)

    The generator ϵ\epsilon in bidegree (1,0)(1,0)

  4. (4)

    Infinitely many elements of the form (92) (see below)

embeds as a bigraded vector space into E1Q{}^{Q}\!E_{1}\otimes\mathbb{R}, because these elements are primitive. The Lie algebra in (1)(1) is at least as large as: the free Lie algebra on {ω5,,ω45}\{\omega^{5},\ldots,\omega^{45}\}, and, the part of the depth-graded motivic Lie algebra generated in depth 11.

7.2. Poincaré series and dimensions

Let

P(s,t)=g,n0dim(T(Ωc[1])g,n)sgtnP(s,t)=\sum_{g,n\geq 0}\dim\left(T(\Omega^{*}_{c}[-1])_{g,n}\right)\,s^{g}t^{n}

denote the Poincaré series of the bigraded vector space T(Ωc[1])T(\Omega^{*}_{c}[-1]) with respect to genus and degree minus genus. Let us define for all k1k\geq 1

f2k+1(t)=t4k+2i=1k1(1+t4i+1).f_{2k+1}(t)=t^{4k+2}\prod_{i=1}^{k-1}(1+t^{4i+1})\ .

Then, for example, we have

f3(t)=t6,f5(t)=(1+t5)t10,f7(t)=(1+t5)(1+t9)t14.f_{3}(t)=t^{6}\ ,\ f_{5}(t)=(1+t^{5})t^{10}\ ,\ f_{7}(t)=(1+t^{5})(1+t^{9})t^{14}\ .

Each polynomial f2k+1(t)f_{2k+1}(t) is the Poincaré series for the graded vector space Ωc(2k+1)[1].\Omega^{*}_{c}(2k+1)[-1]. Since Ωc[1]\Omega^{*}_{c}[-1] is the direct sum of all Ωc(2k+1)[1]\Omega^{*}_{c}(2k+1)[-1] it follows that

P(s,t)=11k1f2k+1(t)s2k+1=1+s3t6+s5(t10+t15)+t12s6+P(s,t)=\frac{1}{1-\sum_{k\geq 1}f_{2k+1}(t)s^{2k+1}}=1+s^{3}t^{6}+s^{5}(t^{10}+t^{15})+t^{12}s^{6}+\ldots

Since f2k+1(1)=0f_{2k+1}(-1)=0 for all k>1k>1, an interesting consequence is that the generating function for the Euler characteristic is P(s,1)=(1s3)1P(s,-1)=(1-s^{3})^{-1}. It follows that the Euler characteristic (with respect to degree) of the genus gg component of T(Ωc[1])T(\Omega^{*}_{c}[-1]) is congruent to 11 if g0(mod3)g\equiv 0\pmod{3} and 0 otherwise.

7.3. Diagonals and degrees; proofs and refinements of Corollaries 1.4, 1.5, and 1.9

The Poincaré series for the tensor algebra T(k1ω4k+1)T(\bigoplus_{k\geq 1}\omega^{4k+1}\mathbb{Q}) featuring in Question 1.17, generated by the classes ω2g1\omega^{2g-1} in bidegree (g,g)(g,g) for g>1g>1 odd, is

P(s,t)=11s3t6s5t10=1s2t41s2t4s3t6P(s,t)=\frac{1}{1-s^{3}t^{6}-s^{5}t^{10}-\ldots}=\frac{1-s^{2}t^{4}}{1-s^{2}t^{4}-s^{3}t^{6}}

The coefficient of snt2ns^{n}t^{2n} in P(t)P(t) is asymptotically αn\alpha^{-n} where α=.7548\alpha=.7548\cdots is the real root of s3+s21=0s^{3}+s^{2}-1=0. Corollary 6.3 and the comments which follow prove that the asymptotic growth of the diagonal gW0Hc2g(𝒜g;)\bigoplus_{g}W_{0}H_{c}^{2g}(\mathcal{A}_{g};\mathbb{Q}) is eventually bounded below by αapproxn\alpha_{\mathrm{approx}}^{-n} where αapprox=0.7551\alpha_{\mathrm{approx}}=0.7551\cdots is the real root of s23+s21++s31=0s^{23}+s^{21}+\ldots+s^{3}-1=0, and is therefore very close to what we would expect if Question 1.17 were true.

In light of the Poincaré–Birkhoff–Witt theorem, by multiplying by any symmetric tensor generated by the elements of Ωc[1]\Omega^{*}_{c}[-1] with the form ω4i1+1ω4ik+1\omega^{4i_{1}+1}\wedge\ldots\wedge\omega^{4i_{k}+1} where k>1k>1, we obtain an additional copy of the diagonal Lie algebra generated by {ω4k+1|k>1}\{\omega^{4k+1}~|~k>1\}, proving Corollary 1.4. By the computation above, this diagonal Lie algebra has exponential growth (see Remark 7.3), proving Corollary 1.5. Here follow two further applications.

7.3.1. Refinement of Corollary 1.5

The dimension of W0Hc2g+k(𝒜g)W_{0}H^{2g+k}_{c}(\mathcal{A}_{g}) grows at least exponentially with gg for all non-negative integers kk except possibly kk in the set

S𝒜={1,2,3,4,6,7,8,10,11,12,15,16,19,20,23,24,28,32,36,40}.S_{\mathcal{A}}=\{1,2,3,4,6,7,8,10,11,12,15,16,19,20,23,24,28,32,36,40\}.

Indeed, by Corollary 1.4, it suffices to show that, for each kS𝒜k\not\in S_{\mathcal{A}}, there is a genus gg such that the bigraded vector space Ωc[1]\Omega_{c}^{*}[-1] is nonzero in genus gg and degree 2g+k2g+k.

Each homogeneous basis element is of the form

ω=ω4k1+1ω4kr+1,\omega=\omega^{4k_{1}+1}\wedge\cdots\wedge\omega^{4k_{r}+1},

where k1<<krk_{1}<\ldots<k_{r}. Then ω\omega is in genus g=2kr+1g=2k_{r}+1 and degree (4k1+1)+(4kr1+1)(4k_{1}+1)\cdots+(4k_{r-1}+1). The claim follows, since the numbers that can be written as a sum of distinct integers that are at least 5 and congruent to 1mod41\mod 4 are

5,9,13,14,17,18,21,22,25,26,27,29,30,31,33,34,35,37,38,39,5,9,13,14,17,18,21,22,25,26,27,29,30,31,33,34,35,37,38,39,

and all integers greater than or equal to 4141.

7.3.2. Refinement of Corollary 1.9

The dimension of H(n2)nk(SLn();)H^{\binom{n}{2}-n-k}(\mathrm{SL}_{n}(\mathbb{Z});\mathbb{Q}) grows at least exponentially for all integers k1k\geq-1 except possibly kk in the set

SSL={1,2,3,6,7,10,11,15,19,23}.S_{\mathrm{SL}}=\{1,2,3,6,7,10,11,15,19,23\}.

We have already noted the bigraded isomorphism

(W0Hc(𝒜))[x]/x2gH(GLg(),Stg),(W_{0}H^{*}_{c}(\mathcal{A}))^{\vee}\otimes_{\mathbb{Q}}\mathbb{Q}[x]/x^{2}\cong\bigoplus_{g}H_{*}(\mathrm{GL}_{g}(\mathbb{Z}),\operatorname{St}_{g}\otimes\mathbb{Q}),

where xx has genus 11 and degree 11. Likewise, we have noted that H(n2)k(SLn();)H^{\binom{n}{2}-k}(\mathrm{SL}_{n}(\mathbb{Z});\mathbb{Q}) contains Hk(GLn(),Stn)H_{k}(\mathrm{GL}_{n}(\mathbb{Z}),\operatorname{St}_{n}\otimes\mathbb{Q}) as a summand. Thus, the dimension of H(n2)nk(SLn();)H^{\binom{n}{2}-n-k}(\mathrm{SL}_{n}(\mathbb{Z});\mathbb{Q}) grows exponentially with nn unless both kk and k+1k+1 are in S𝒜S_{\mathcal{A}}.

7.4. Polynomials in the wheel homology classes

The cohomology classes defined above are only defined over the real numbers, because their pairing with rational homology cycles are given by period integrals which include odd values of the zeta function. However, it was shown in [BS24] that the wheel classes give explicit rational homology classes: for all odd g>1g>1 there exists an explicit non-zero locally finite homology class:

(89) [τWg]H2gBM(Pg/GLg();).[\tau_{W_{g}}]\in H^{BM}_{2g}(P_{g}/\mathrm{GL}_{g}(\mathbb{Z});\mathbb{Q})\ .

A corollary of the existence of the Hopf algebra structure on the E1E^{1}-page of the homological Quillen spectral sequence implies the following:

Theorem 7.4.

There is an injective map of commutative bigraded algebras

(90) [W3,W5,,Wg,][x]/x2E1Q\mathbb{Q}[W_{3},W_{5},\ldots,W_{g},\ldots]\otimes\mathbb{Q}[x]/x^{2}\longrightarrow{}^{Q}\!E^{1}

where WgW_{g} denotes the wheel class (89) in odd genus gg and degree 2g2g, and where the element xx is in degree 11 and genus 11 and maps to ee. The bigrading on the left hand side is by genus and degree minus genus.

Proof.

By their definition, the wheel classes (89) factor through the graph spectral sequence E1GE1Q{}^{G}\!E^{1}\rightarrow{}^{Q}\!E^{1}. They are primitive in E1Q{}^{Q}\!E^{1} because this is true a fortiori in the graph complex, by Lemma 6.1. The element xx, which represents a one-vertex, one-edge loop in E1,01G{}^{G}\!E^{1}_{1,0}, is primitive for reasons of degree.

The homological E1E^{1}-page is graded-commutative but not primitively generated. The Milnor–Moore theorem ([MM65, Theorem 5.18]) implies that the canonical map

Sym(Prim(E,1Q))=𝒰(Prim(E,1Q))E,1Q\mathrm{Sym}(\operatorname{Prim}({}^{Q}\!E^{1}_{*,*}))=\mathcal{U}(\operatorname{Prim}({}^{Q}\!E^{1}_{*,*}))\to{}^{Q}\!E^{1}_{*,*}

is an isomorphism onto the subalgebra generated by Prim(E,1Q)E,1Q\operatorname{Prim}({}^{Q}\!E^{1}_{*,*})\subset{}^{Q}\!E^{1}_{*,*}, and the domain contains [W3,W5,][x]/x2\mathbb{Q}[W_{3},W_{5},\ldots]\otimes\mathbb{Q}[x]/x^{2} as a subalgebra. ∎

It was shown in [BS24] that the wheel classes WgW_{g} pair non-trivially with the primitive canonical forms ω2g1\omega^{2g-1}. It follows that the dual to (90), tensored with \mathbb{R}, is the map

W0Hc(𝒜;)[ω5,ω9,,ω2g1,].W_{0}H_{c}^{*}(\mathcal{A};\mathbb{Q})\otimes_{\mathbb{Q}}\mathbb{R}\longrightarrow\mathbb{R}[\omega^{5},\omega^{9},\ldots,\omega^{2g-1},\ldots]\ .

which sends all other primitives to zero. In fact, one may replace \mathbb{R} in the previous map with the \mathbb{Q}-algebra generated by odd zeta values ζ(2n+1)\zeta(2n+1), for n1.n\geq 1.

7.5. The canonical spectral sequence

Recall the map

T(Ωc[1])W0Hc(𝒜;)T(\Omega^{*}_{c}[-1]\otimes\mathbb{R})\to W_{0}H^{*}_{c}(\mathcal{A};\mathbb{R})

which is injective on Ωc[1]\Omega^{*}_{c}[-1] by Theorem 4.5. Here we construct a spectral sequence whose E1E_{1}-page is T(Ωc[1])T(\Omega^{*}_{c}[-1]) and compare it to the cohomological Quillen spectral sequence. Consider the graded exterior algebra P\wedge P^{*} on the graded vector space P=k1β4k+1P^{*}=\bigoplus_{k\geq 1}\mathbb{Q}\beta^{4k+1} with generators β4k+1\beta^{4k+1} in degree 4k+14k+1. Endowed with the zero differential, P\wedge P^{*} defines a connected differential graded algebra. The bar construction B(P)B(\wedge P^{*}) is a graded commutative Hopf algebra over \mathbb{Q} generated by symbols [p1||pn][p_{1}|\ldots|p_{n}] in degree deg(p1)++deg(pn)+n\deg(p_{1})+\ldots+\deg(p_{n})+n where deg(pi)>0\deg(p_{i})>0 (note that a more standard convention is to have a minus before the nn; this is not the case here), where pip_{i} are homogeneous generators of P\wedge P^{*}, and is isomorphic as a vector space to the tensor algebra T((P+))T(\wedge(P^{*}_{+})), where P+P^{*}_{+} denotes the part of PP^{*} in positive degree. In addition it is equipped with the (graded-commutative) signed shuffle product, and the deconcatenation coproduct

Δ[p1||pn]=i=1n[p1||pi][pi+1||pn].\Delta[p_{1}|\ldots|p_{n}]=\sum_{i=1}^{n}[p_{1}|\ldots|p_{i}]\otimes[p_{i+1}|\ldots|p_{n}]\ .

In addition, there is an internal differential

(91) dI:B(P)\displaystyle d_{I}\colon B(\wedge P^{*}) \displaystyle\longrightarrow B(P)\displaystyle B(\wedge P^{*})
dI([p1||pn])\displaystyle d_{I}([p_{1}|\ldots|p_{n}]) =\displaystyle= i=1n1(1)i[sp1||spi1|spipi+1|pi+2||pn]\displaystyle\sum_{i=1}^{n-1}(-1)^{i}[sp_{1}|\ldots|sp_{i-1}|sp_{i}\wedge p_{i+1}|p_{i+2}|\ldots|p_{n}]

where s:PPs\colon\wedge P^{*}\rightarrow\wedge P^{*} is the linear map of graded vector spaces which multiplies by (1)n(-1)^{n} in degree nn. The differential dId_{I} has degree 1-1 (owing to the plus sign in our convention for the degree of bar elements) and satisfies dI2=0d_{I}^{2}=0. It is compatible with the Hopf algebra structures.

Define an increasing filtration GG on P\wedge P^{*} as follows: Gg(P)G_{g}(\wedge P^{*}) is spanned by elements ω4i1+1ω4ik+1\omega^{4i_{1}+1}\wedge\ldots\wedge\omega^{4i_{k}+1} such that i1<<iki_{1}<\ldots<i_{k} and 4ik+12g14i_{k}+1\leq 2g-1. It defines a filtration G(P)G(\wedge P^{*}) of differential graded algebras. In fact, it follows from the definition that GgGhGmax{g,h}G_{g}\wedge G_{h}\subset G_{\max\{g,h\}}. The filtration GG induces a filtration on B(P)B(\wedge P^{*}), which we also denote by GG. It is a filtration of graded Hopf algebras, which is respected by the differential dId_{I}.

Proposition 7.5.

The filtered complex (B(P),dI)(B(\wedge P^{*}),d_{I}) with filtration GG defines a spectral sequence Es,trc{}^{c}\!E^{r}_{s,t} of commutative bigraded Hopf algebras such that

Es,t1c=GrsGBs+t(P).{}^{c}\!E^{1}_{s,t}=\operatorname{Gr}_{s}^{G}B_{s+t}(\wedge P^{*})\ .

This spectral sequence converges to the bigraded Hopf algebra GrGSym(P+[1])\operatorname{Gr}^{G}\mathrm{Sym}(P_{+}^{*}[-1]) which is isomorphic to the polynomial ring in primitive generators β4k+1\beta^{4k+1} in degree 4k+24k+2 and genus 2k+12k+1. Furthermore, the differentials d1,d2d_{1},d_{2} vanish, and dr[β4k+1]=0d_{r}[\beta^{4k+1}]=0 for all k,rk,r.

Proof.

The spectral sequence Es,t1c{}^{c}\!E^{1}_{s,t} defined by the filtration GG on (B(P),dI)(B(\wedge P^{*}),d_{I}) has E1E^{1}-page isomorphic to

Es,t1c=Hs+t(GrsGB(P)){}^{c}\!E^{1}_{s,t}=H_{s+t}(\operatorname{Gr}_{s}^{G}B(\wedge P^{*}))

Since the differential dId_{I} strictly decreases the genus, it is identically zero on the associated graded of B(P)B(\wedge P^{*}), and hence Hs+t(GrsGB(P))=GrsGBs+t(P)H_{s+t}(\operatorname{Gr}_{s}^{G}B(\wedge P^{*}))=\operatorname{Gr}_{s}^{G}B_{s+t}(\wedge P^{*}). It follows from the Koszul duality between the symmetric and exterior algebras ([LV12], Proposition 3.48 and Theorem 3.44) that there is an isomorphism of graded commutative Hopf algebras

H(B(P),dI)Sym(P+[1]).H_{*}\left(B(\wedge P^{*}),d_{I}\right)\cong\mathrm{Sym}(P_{+}^{*}[-1])\ .

The spectral sequence therefore converges to GrGH(B(P),dI)GrGSym(P+[1])\operatorname{Gr}^{G}H(B(\wedge P^{*}),d_{I})\cong\operatorname{Gr}^{G}\mathrm{Sym}(P_{+}^{*}[-1]), as bigraded Hopf algebras. The fact that the differential drd_{r} annihilates β4k+1\beta^{4k+1} is clear from the definition of dId_{I}, which acts trivially on [β4k+1][\beta^{4k+1}]. The fact that d1,d2d_{1},d_{2} vanish follows from the definition of dId_{I}, and the fact that the map [β4a+1|β4b+1]β4a+1β4b+1[\beta^{4a+1}|\beta^{4b+1}]\mapsto\beta^{4a+1}\wedge\beta^{4b+1} sends a term of genus 2(a+b)+22(a+b)+2 to one of genus 2max{a,b}+22\max\{a,b\}+2, and therefore decreases the genus by at least 33. ∎

By identifying PP^{*} with the graded dual of Ωc\Omega^{*}_{c}, we may interpret B(P)B(\wedge P^{*}) as the graded dual of the tensor Hopf algebra T(Ωc[1])T(\Omega^{*}_{c}[-1]). The proposition therefore defines by duality a spectral sequence on T(Ωc[1])T(\Omega^{*}_{c}[-1]), which we call the canonical spectral sequence, denoted by Ers,tc{}^{c}\!E^{s,t}_{r}. The differentials in this spectral sequence vanish on elements [ω4k+1][\omega^{4k+1}] and, for r=2min{a,b}+1r=2\min\{a,b\}+1, send [ω4a+1ω4b+1][\omega^{4a+1}\wedge\omega^{4b+1}] to the commutator [ω4a+1,ω4b+1]=[ω4a+1|ω4b+1][ω4b+1|ω4a+1].[\omega^{4a+1},\omega^{4b+1}]=[\omega^{4a+1}|\omega^{4b+1}]-[\omega^{4b+1}|\omega^{4a+1}].

We expect that the map in Theorem 4.5 may be promoted (possibly after rescaling the action of the differentials) to a map of spectral sequences

Er,cEr,Q{}^{c}\!E^{*,*}_{r}\otimes_{\mathbb{Q}}\mathbb{R}\longrightarrow{}^{Q}\!E_{r}^{*,*}\otimes_{\mathbb{Q}}\mathbb{R}\

which induces an isomorphism on their abutments (which are formally isomorphic, by the previous proposition). This provides yet more evidence of a different kind for Conjecture 1.16.

7.6. Illustration

Table 1 depicts T(Ωc[1])T(\Omega^{*}_{c}[-1]). The entries that are known to be isomorphic to W0Hc(𝒜;)W_{0}H^{*}_{c}(\mathcal{A};\mathbb{Q}) are highlighted, in particular for g7g\leq 7. Blank entries vanish for dimension reasons. Entries in low genus follow from computer calculations of [EVGS13] and [DSEVKM19].

There are two infinite ranges in which the cohomology of W0Hc(𝒜;)W_{0}H^{*}_{c}(\mathcal{A};\mathbb{Q}) has been completely determined. The zero entries in the bottom three rows follow from [Gun00, BPS23, BMP+24], which imply that

W0Hcg+n(𝒜g;)=0 for 0n2,g1.W_{0}H^{g+n}_{c}(\mathcal{A}_{g};\mathbb{Q})=0\quad\hbox{ for }0\leq n\leq 2,g\geq 1\ .

The bottom two rows also follow from [LS76] and [CP17]. A conjecture of Church–Farb–Putnam implies that W0Hcn(𝒜g;)W_{0}H^{n}_{c}(\mathcal{A}_{g};\mathbb{Q}) vanishes for n<2gn<2g (below the diagonal line n=2gn=2g). The entries in high degrees follow from [Bro23, §14.5], [BBC+24] and imply that for all g>1g>1 odd

W0Hcn(𝒜g;)Ωcn(g)[1] and W0Hcn1(𝒜g1;)=0W_{0}H_{c}^{n}(\mathcal{A}_{g};\mathbb{Q})\cong\Omega_{c}^{n}(g)[-1]\quad\hbox{ and }\quad W_{0}H_{c}^{n-1}(\mathcal{A}_{g-1};\mathbb{Q})=0

for ndgκ(g)n\geq d_{g}-\kappa(g), where dgd_{g} is the dimension of 𝒜g\mathcal{A}_{g} and κ(g)\kappa(g) is the stable range for the cohomology of the general linear group (which is currently known to be κ(g)g1\kappa(g)\leq g-1 by [LS19]).

45 𝟎{\color[rgb]{0,0.6,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.6,0}\pgfsys@color@cmyk@stroke{0.64}{0}{0.95}{0.40}\pgfsys@color@cmyk@fill{0.64}{0}{0.95}{0.40}\bm{0}}
44 𝟎{\color[rgb]{0,0.6,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.6,0}\pgfsys@color@cmyk@stroke{0.64}{0}{0.95}{0.40}\pgfsys@color@cmyk@fill{0.64}{0}{0.95}{0.40}\bm{0}}
43 0
42 0
41 0
40 0
39 0
38 0
37 0
36 [𝝎𝟓𝝎𝟗𝝎𝟏𝟑𝝎𝟏𝟕]{\color[rgb]{0,0.6,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.6,0}\pgfsys@color@cmyk@stroke{0.64}{0}{0.95}{0.40}\pgfsys@color@cmyk@fill{0.64}{0}{0.95}{0.40}\bm{[\omega^{5}\!\wedge\!\omega^{9}\!\wedge\!\omega^{13}\!\wedge\!\omega^{17}]}} 0
35 𝟎{\color[rgb]{0,0.6,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.6,0}\pgfsys@color@cmyk@stroke{0.64}{0}{0.95}{0.40}\pgfsys@color@cmyk@fill{0.64}{0}{0.95}{0.40}\bm{0}} 0
34 𝟎{\color[rgb]{0,0.6,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.6,0}\pgfsys@color@cmyk@stroke{0.64}{0}{0.95}{0.40}\pgfsys@color@cmyk@fill{0.64}{0}{0.95}{0.40}\bm{0}} 0
33 𝟎{\color[rgb]{0,0.6,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.6,0}\pgfsys@color@cmyk@stroke{0.64}{0}{0.95}{0.40}\pgfsys@color@cmyk@fill{0.64}{0}{0.95}{0.40}\bm{0}} 0
32 𝟎{\color[rgb]{0,0.6,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.6,0}\pgfsys@color@cmyk@stroke{0.64}{0}{0.95}{0.40}\pgfsys@color@cmyk@fill{0.64}{0}{0.95}{0.40}\bm{0}} 0
31 [𝝎𝟗𝝎𝟏𝟑𝝎𝟏𝟕]{\color[rgb]{0,0.6,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.6,0}\pgfsys@color@cmyk@stroke{0.64}{0}{0.95}{0.40}\pgfsys@color@cmyk@fill{0.64}{0}{0.95}{0.40}\bm{[\omega^{9}\!\wedge\!\omega^{13}\!\wedge\!\omega^{17}]}} 0
30 𝟎{\color[rgb]{0,0.6,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.6,0}\pgfsys@color@cmyk@stroke{0.64}{0}{0.95}{0.40}\pgfsys@color@cmyk@fill{0.64}{0}{0.95}{0.40}\bm{0}} 0
29 𝟎{\color[rgb]{0,0.6,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.6,0}\pgfsys@color@cmyk@stroke{0.64}{0}{0.95}{0.40}\pgfsys@color@cmyk@fill{0.64}{0}{0.95}{0.40}\bm{0}} 0
28 𝟎{\color[rgb]{0,0.6,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.6,0}\pgfsys@color@cmyk@stroke{0.64}{0}{0.95}{0.40}\pgfsys@color@cmyk@fill{0.64}{0}{0.95}{0.40}\bm{0}} 𝟎{\color[rgb]{0,0.6,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.6,0}\pgfsys@color@cmyk@stroke{0.64}{0}{0.95}{0.40}\pgfsys@color@cmyk@fill{0.64}{0}{0.95}{0.40}\bm{0}} 0
27 𝟎{\color[rgb]{0,0.6,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.6,0}\pgfsys@color@cmyk@stroke{0.64}{0}{0.95}{0.40}\pgfsys@color@cmyk@fill{0.64}{0}{0.95}{0.40}\bm{0}} [ω5ω13ω17][\omega^{5}\!\wedge\!\omega^{13}\!\wedge\!\omega^{17}] 0
26 0 0 0
25 0 0 0
24 0 0 [ω5|ω5ω9ω13][ω5ω9ω13|ω5]\begin{subarray}{c}[\omega^{5}|\omega^{5}\wedge\omega^{9}\wedge\omega^{13}]\\ {\color[rgb]{1,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{1,0,0}[\omega^{5}\wedge\omega^{9}\wedge\omega^{13}|\omega^{5}]}\end{subarray}
23 0 [ω5ω9ω17][\omega^{5}\!\wedge\!\omega^{9}\!\wedge\!\omega^{17}] 0
22 0 [ω13ω17][\omega^{13}\!\wedge\!\omega^{17}] 0
21 [𝝎𝟓𝝎𝟗𝝎𝟏𝟑]{\color[rgb]{0,0.6,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.6,0}\pgfsys@color@cmyk@stroke{0.64}{0}{0.95}{0.40}\pgfsys@color@cmyk@fill{0.64}{0}{0.95}{0.40}\bm{[\omega^{5}\!\wedge\!\omega^{9}\!\wedge\!\omega^{13}]}} 0 0 0
20 𝟎{\color[rgb]{0,0.6,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.6,0}\pgfsys@color@cmyk@stroke{0.64}{0}{0.95}{0.40}\pgfsys@color@cmyk@fill{0.64}{0}{0.95}{0.40}\bm{0}} 0 0 [ω5ω9|ω5ω9]{\color[rgb]{1,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{1,0,0}[\omega^{5}\!\wedge\!\omega^{9}|\omega^{5}\!\wedge\!\omega^{9}]}
19 𝟎{\color[rgb]{0,0.6,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.6,0}\pgfsys@color@cmyk@stroke{0.64}{0}{0.95}{0.40}\pgfsys@color@cmyk@fill{0.64}{0}{0.95}{0.40}\bm{0}} 0 0 [ω5|ω9ω13][ω9ω13|ω5]\begin{subarray}{c}[\omega^{5}|\omega^{9}\wedge\omega^{13}]\\ {\color[rgb]{1,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{1,0,0}[\omega^{9}\wedge\omega^{13}|\omega^{5}]}\end{subarray}
18 𝟎{\color[rgb]{0,0.6,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.6,0}\pgfsys@color@cmyk@stroke{0.64}{0}{0.95}{0.40}\pgfsys@color@cmyk@fill{0.64}{0}{0.95}{0.40}\bm{0}} 0 [ω9ω17][\omega^{9}\!\wedge\!\omega^{17}] 0
17 𝟎{\color[rgb]{0,0.6,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.6,0}\pgfsys@color@cmyk@stroke{0.64}{0}{0.95}{0.40}\pgfsys@color@cmyk@fill{0.64}{0}{0.95}{0.40}\bm{0}} 0 0 0
16 [𝝎𝟗𝝎𝟏𝟑]{\color[rgb]{0,0.6,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.6,0}\pgfsys@color@cmyk@stroke{0.64}{0}{0.95}{0.40}\pgfsys@color@cmyk@fill{0.64}{0}{0.95}{0.40}\bm{[\omega^{9}\!\wedge\!\omega^{13}]}} 0 0 0
15 𝟎{\color[rgb]{0,0.6,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.6,0}\pgfsys@color@cmyk@stroke{0.64}{0}{0.95}{0.40}\pgfsys@color@cmyk@fill{0.64}{0}{0.95}{0.40}\bm{0}} 𝟎{\color[rgb]{0,0.6,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.6,0}\pgfsys@color@cmyk@stroke{0.64}{0}{0.95}{0.40}\pgfsys@color@cmyk@fill{0.64}{0}{0.95}{0.40}\bm{0}} 0 0 [ω9|ω5ω9][ω5ω9|ω9],[ω5|ω5ω13][ω5ω13|ω5]\begin{subarray}{c}[\omega^{9}|\omega^{5}\wedge\omega^{9}]\\ [\omega^{5}\wedge\omega^{9}|\omega^{9}]\end{subarray},\begin{subarray}{c}[\omega^{5}|\omega^{5}\wedge\omega^{13}]\\ {\color[rgb]{1,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{1,0,0}[\omega^{5}\wedge\omega^{13}|\omega^{5}]}\end{subarray}
14 𝟎{\color[rgb]{0,0.6,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.6,0}\pgfsys@color@cmyk@stroke{0.64}{0}{0.95}{0.40}\pgfsys@color@cmyk@fill{0.64}{0}{0.95}{0.40}\bm{0}} 𝟎{\color[rgb]{0,0.5,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.5,1}\pgfsys@color@cmyk@stroke{1}{0.50}{0}{0}\pgfsys@color@cmyk@fill{1}{0.50}{0}{0}\bm{0}} 0 [ω5ω17][\omega^{5}\!\wedge\!\omega^{17}] 0
13 𝟎{\color[rgb]{0,0.5,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.5,1}\pgfsys@color@cmyk@stroke{1}{0.50}{0}{0}\pgfsys@color@cmyk@fill{1}{0.50}{0}{0}\bm{0}} 𝟎{\color[rgb]{0,0.5,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.5,1}\pgfsys@color@cmyk@stroke{1}{0.50}{0}{0}\pgfsys@color@cmyk@fill{1}{0.50}{0}{0}\bm{0}} [ω5|ω5ω9][ω5ω9|ω5]\begin{subarray}{c}[\omega^{5}|\omega^{5}\wedge\omega^{9}]\\ [\omega^{5}\wedge\omega^{9}|\omega^{5}]\end{subarray} 0 0
12 𝟎{\color[rgb]{0,0.5,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.5,1}\pgfsys@color@cmyk@stroke{1}{0.50}{0}{0}\pgfsys@color@cmyk@fill{1}{0.50}{0}{0}\bm{0}} [𝝎𝟓𝝎𝟏𝟑]{\color[rgb]{0,0.5,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.5,1}\pgfsys@color@cmyk@stroke{1}{0.50}{0}{0}\pgfsys@color@cmyk@fill{1}{0.50}{0}{0}\bm{[\omega^{5}\!\wedge\!\omega^{13}]}} 0 0 0
11 𝟎{\color[rgb]{0,0.5,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.5,1}\pgfsys@color@cmyk@stroke{1}{0.50}{0}{0}\pgfsys@color@cmyk@fill{1}{0.50}{0}{0}\bm{0}} 𝟎{\color[rgb]{0,0.5,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.5,1}\pgfsys@color@cmyk@stroke{1}{0.50}{0}{0}\pgfsys@color@cmyk@fill{1}{0.50}{0}{0}\bm{0}} 0 0 0
10 [𝝎𝟓𝝎𝟗]{\color[rgb]{0,0.6,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.6,0}\pgfsys@color@cmyk@stroke{0.64}{0}{0.95}{0.40}\pgfsys@color@cmyk@fill{0.64}{0}{0.95}{0.40}\bm{[\omega^{5}\!\wedge\!\omega^{9}]}} 𝟎{\color[rgb]{0,0.5,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.5,1}\pgfsys@color@cmyk@stroke{1}{0.50}{0}{0}\pgfsys@color@cmyk@fill{1}{0.50}{0}{0}\bm{0}} 𝟎{\color[rgb]{0,0.5,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.5,1}\pgfsys@color@cmyk@stroke{1}{0.50}{0}{0}\pgfsys@color@cmyk@fill{1}{0.50}{0}{0}\bm{0}} 0 0 [ω9|ω9][ω5|ω13],[ω13|ω5]\begin{subarray}{c}[\omega^{9}|\omega^{9}]\\ [\omega^{5}|\omega^{13}]\ ,\ [\omega^{13}|\omega^{5}]\end{subarray}
9 𝟎{\color[rgb]{0,0.6,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.6,0}\pgfsys@color@cmyk@stroke{0.64}{0}{0.95}{0.40}\pgfsys@color@cmyk@fill{0.64}{0}{0.95}{0.40}\bm{0}} 𝟎{\color[rgb]{0,0.5,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.5,1}\pgfsys@color@cmyk@stroke{1}{0.50}{0}{0}\pgfsys@color@cmyk@fill{1}{0.50}{0}{0}\bm{0}} 𝟎{\color[rgb]{0,0.5,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.5,1}\pgfsys@color@cmyk@stroke{1}{0.50}{0}{0}\pgfsys@color@cmyk@fill{1}{0.50}{0}{0}\bm{0}} 0 [ω17][ω5|ω5|ω5]\begin{subarray}{c}[\omega^{17}]\\ [\omega^{5}|\omega^{5}|\omega^{5}]\end{subarray} 0
8 𝟎{\color[rgb]{0,0.6,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.6,0}\pgfsys@color@cmyk@stroke{0.64}{0}{0.95}{0.40}\pgfsys@color@cmyk@fill{0.64}{0}{0.95}{0.40}\bm{0}} 𝟎{\color[rgb]{0,0.5,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.5,1}\pgfsys@color@cmyk@stroke{1}{0.50}{0}{0}\pgfsys@color@cmyk@fill{1}{0.50}{0}{0}\bm{0}} 𝟎{\color[rgb]{0,0.5,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.5,1}\pgfsys@color@cmyk@stroke{1}{0.50}{0}{0}\pgfsys@color@cmyk@fill{1}{0.50}{0}{0}\bm{0}} [ω5|ω9][ω9|ω5]\begin{subarray}{c}[\omega^{5}|\omega^{9}]\\ [\omega^{9}|\omega^{5}]\end{subarray} 0 0
7 𝟎{\color[rgb]{0,0.6,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.6,0}\pgfsys@color@cmyk@stroke{0.64}{0}{0.95}{0.40}\pgfsys@color@cmyk@fill{0.64}{0}{0.95}{0.40}\bm{0}} 𝟎{\color[rgb]{0,0.5,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.5,1}\pgfsys@color@cmyk@stroke{1}{0.50}{0}{0}\pgfsys@color@cmyk@fill{1}{0.50}{0}{0}\bm{0}} [𝝎𝟏𝟑]{\color[rgb]{0,0.5,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.5,1}\pgfsys@color@cmyk@stroke{1}{0.50}{0}{0}\pgfsys@color@cmyk@fill{1}{0.50}{0}{0}\bm{[\omega^{13}]}} 0 0 0
6 𝟎{\color[rgb]{0,0.6,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.6,0}\pgfsys@color@cmyk@stroke{0.64}{0}{0.95}{0.40}\pgfsys@color@cmyk@fill{0.64}{0}{0.95}{0.40}\bm{0}} 𝟎{\color[rgb]{0,0.6,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.6,0}\pgfsys@color@cmyk@stroke{0.64}{0}{0.95}{0.40}\pgfsys@color@cmyk@fill{0.64}{0}{0.95}{0.40}\bm{0}} [𝝎𝟓|𝝎𝟓]{\color[rgb]{0,0.5,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.5,1}\pgfsys@color@cmyk@stroke{1}{0.50}{0}{0}\pgfsys@color@cmyk@fill{1}{0.50}{0}{0}\bm{[\omega^{5}|\omega^{5}]}} 𝟎{\color[rgb]{0,0.5,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.5,1}\pgfsys@color@cmyk@stroke{1}{0.50}{0}{0}\pgfsys@color@cmyk@fill{1}{0.50}{0}{0}\bm{0}} 0 0 0
5 𝟎{\color[rgb]{0,0.6,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.6,0}\pgfsys@color@cmyk@stroke{0.64}{0}{0.95}{0.40}\pgfsys@color@cmyk@fill{0.64}{0}{0.95}{0.40}\bm{0}} [𝝎𝟗]{\color[rgb]{0,0.5,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.5,1}\pgfsys@color@cmyk@stroke{1}{0.50}{0}{0}\pgfsys@color@cmyk@fill{1}{0.50}{0}{0}\bm{[\omega^{9}]}} 𝟎{\color[rgb]{0,0.5,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.5,1}\pgfsys@color@cmyk@stroke{1}{0.50}{0}{0}\pgfsys@color@cmyk@fill{1}{0.50}{0}{0}\bm{0}} 𝟎{\color[rgb]{0,0.5,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.5,1}\pgfsys@color@cmyk@stroke{1}{0.50}{0}{0}\pgfsys@color@cmyk@fill{1}{0.50}{0}{0}\bm{0}} 0 0 0
4 𝟎{\color[rgb]{0,0.5,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.5,1}\pgfsys@color@cmyk@stroke{1}{0.50}{0}{0}\pgfsys@color@cmyk@fill{1}{0.50}{0}{0}\bm{0}} 𝟎{\color[rgb]{0,0.5,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.5,1}\pgfsys@color@cmyk@stroke{1}{0.50}{0}{0}\pgfsys@color@cmyk@fill{1}{0.50}{0}{0}\bm{0}} 𝟎{\color[rgb]{0,0.5,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.5,1}\pgfsys@color@cmyk@stroke{1}{0.50}{0}{0}\pgfsys@color@cmyk@fill{1}{0.50}{0}{0}\bm{0}} 𝟎{\color[rgb]{0,0.5,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.5,1}\pgfsys@color@cmyk@stroke{1}{0.50}{0}{0}\pgfsys@color@cmyk@fill{1}{0.50}{0}{0}\bm{0}} 𝟎{\color[rgb]{0,0.5,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.5,1}\pgfsys@color@cmyk@stroke{1}{0.50}{0}{0}\pgfsys@color@cmyk@fill{1}{0.50}{0}{0}\bm{0}} 0 0
3 [𝝎𝟓]{\color[rgb]{0,0.6,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.6,0}\pgfsys@color@cmyk@stroke{0.64}{0}{0.95}{0.40}\pgfsys@color@cmyk@fill{0.64}{0}{0.95}{0.40}\bm{[\omega^{5}]}} 𝟎{\color[rgb]{0,0.5,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.5,1}\pgfsys@color@cmyk@stroke{1}{0.50}{0}{0}\pgfsys@color@cmyk@fill{1}{0.50}{0}{0}\bm{0}} 𝟎{\color[rgb]{0,0.5,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.5,1}\pgfsys@color@cmyk@stroke{1}{0.50}{0}{0}\pgfsys@color@cmyk@fill{1}{0.50}{0}{0}\bm{0}} 𝟎{\color[rgb]{0,0.5,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.5,1}\pgfsys@color@cmyk@stroke{1}{0.50}{0}{0}\pgfsys@color@cmyk@fill{1}{0.50}{0}{0}\bm{0}} 𝟎{\color[rgb]{0,0.5,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.5,1}\pgfsys@color@cmyk@stroke{1}{0.50}{0}{0}\pgfsys@color@cmyk@fill{1}{0.50}{0}{0}\bm{0}} 𝟎{\color[rgb]{0,0.5,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.5,1}\pgfsys@color@cmyk@stroke{1}{0.50}{0}{0}\pgfsys@color@cmyk@fill{1}{0.50}{0}{0}\bm{0}} 0 0
2 𝟎{\color[rgb]{.5,0,.5}\definecolor[named]{pgfstrokecolor}{rgb}{.5,0,.5}\bm{0}} 𝟎{\color[rgb]{.5,0,.5}\definecolor[named]{pgfstrokecolor}{rgb}{.5,0,.5}\bm{0}} 𝟎{\color[rgb]{.5,0,.5}\definecolor[named]{pgfstrokecolor}{rgb}{.5,0,.5}\bm{0}} 𝟎{\color[rgb]{.5,0,.5}\definecolor[named]{pgfstrokecolor}{rgb}{.5,0,.5}\bm{0}} 𝟎{\color[rgb]{.5,0,.5}\definecolor[named]{pgfstrokecolor}{rgb}{.5,0,.5}\bm{0}} 𝟎{\color[rgb]{.5,0,.5}\definecolor[named]{pgfstrokecolor}{rgb}{.5,0,.5}\bm{0}} 𝟎{\color[rgb]{.5,0,.5}\definecolor[named]{pgfstrokecolor}{rgb}{.5,0,.5}\bm{0}} 𝟎{\color[rgb]{.5,0,.5}\definecolor[named]{pgfstrokecolor}{rgb}{.5,0,.5}\bm{0}}
1 𝟎{\color[rgb]{.5,0,.5}\definecolor[named]{pgfstrokecolor}{rgb}{.5,0,.5}\bm{0}} 𝟎{\color[rgb]{.5,0,.5}\definecolor[named]{pgfstrokecolor}{rgb}{.5,0,.5}\bm{0}} 𝟎{\color[rgb]{.5,0,.5}\definecolor[named]{pgfstrokecolor}{rgb}{.5,0,.5}\bm{0}} 𝟎{\color[rgb]{.5,0,.5}\definecolor[named]{pgfstrokecolor}{rgb}{.5,0,.5}\bm{0}} 𝟎{\color[rgb]{.5,0,.5}\definecolor[named]{pgfstrokecolor}{rgb}{.5,0,.5}\bm{0}} 𝟎{\color[rgb]{.5,0,.5}\definecolor[named]{pgfstrokecolor}{rgb}{.5,0,.5}\bm{0}} 𝟎{\color[rgb]{.5,0,.5}\definecolor[named]{pgfstrokecolor}{rgb}{.5,0,.5}\bm{0}} 𝟎{\color[rgb]{.5,0,.5}\definecolor[named]{pgfstrokecolor}{rgb}{.5,0,.5}\bm{0}} 𝟎{\color[rgb]{.5,0,.5}\definecolor[named]{pgfstrokecolor}{rgb}{.5,0,.5}\bm{0}}
0 𝟏{\color[rgb]{.5,0,.5}\definecolor[named]{pgfstrokecolor}{rgb}{.5,0,.5}\bm{1}} 𝟎{\color[rgb]{.5,0,.5}\definecolor[named]{pgfstrokecolor}{rgb}{.5,0,.5}\bm{0}} 𝟎{\color[rgb]{.5,0,.5}\definecolor[named]{pgfstrokecolor}{rgb}{.5,0,.5}\bm{0}} 𝟎{\color[rgb]{.5,0,.5}\definecolor[named]{pgfstrokecolor}{rgb}{.5,0,.5}\bm{0}} 𝟎{\color[rgb]{.5,0,.5}\definecolor[named]{pgfstrokecolor}{rgb}{.5,0,.5}\bm{0}} 𝟎{\color[rgb]{.5,0,.5}\definecolor[named]{pgfstrokecolor}{rgb}{.5,0,.5}\bm{0}} 𝟎{\color[rgb]{.5,0,.5}\definecolor[named]{pgfstrokecolor}{rgb}{.5,0,.5}\bm{0}} 𝟎{\color[rgb]{.5,0,.5}\definecolor[named]{pgfstrokecolor}{rgb}{.5,0,.5}\bm{0}} 𝟎{\color[rgb]{.5,0,.5}\definecolor[named]{pgfstrokecolor}{rgb}{.5,0,.5}\bm{0}} 𝟎{\color[rgb]{.5,0,.5}\definecolor[named]{pgfstrokecolor}{rgb}{.5,0,.5}\bm{0}} 𝟎{\color[rgb]{.5,0,.5}\definecolor[named]{pgfstrokecolor}{rgb}{.5,0,.5}\bm{0}}
0 1 2 3 4 5 6 7 8 9 10
Table 1. Without the red entries, a table showing generators for a largest known subspace of W0Hc(𝒜;)W_{0}H^{*}_{c}(\mathcal{A};\mathbb{R}) in genus up to 1010, deduced from Theorems 1.2 and 1.3, and the discussion in Section 7.6. The (s,t)(s,t) entry shows linearly independent elements of W0Hcs+t(𝒜s;)W_{0}H^{s+t}_{c}(\mathcal{A}_{s};\mathbb{R}). An entry (s,t)(s,t) is highlighted in green, blue, or purple if it is known to be all of W0Hcs+t(𝒜s;)W_{0}H^{s+t}_{c}(\mathcal{A}_{s};\mathbb{R}); the color coding is given in Section 7.6. The red entries are additional elements of T(Ωc[1])T(\Omega^{*}_{c}[-1]) appearing as elements of E1E_{1} of the canonical spectral sequence (Section 7.5), and conjecturally (Conjecture 1.16) new linearly independent generators of W0Hc(𝒜;)W_{0}H^{*}_{c}(\mathcal{A};\mathbb{R}). Thus, the table in its entirety shows E1c{}^{c}\!E_{1} and verifies Conjecture 1.16 for all g9g\leq 9.

7.6.1. Application of the Quillen spectral sequence

Studying the Quillen spectral sequence further allows us to deduce some nonvanishing classes in W0H(𝒜;)W_{0}H^{*}(\mathcal{A};\mathbb{R}) and verifying Conjecture 1.16 up to g=9g=9. Recall that the cohomological Quillen spectral sequence abuts to a free polynomial algebra with generators in bidegree (2k+1,2k+1)(2k+1,2k+1), for positive integers kk. By inspection of Table 1, we observe that [ω5][\omega^{5}] and [ω9][\omega^{9}] (and in fact, [ω13][\omega^{13}], since a class in genus 7 and degree 14 must appear in the abutment of the Quillen spectral sequence) are annihilated by the differentials drQ{}^{Q}d_{r} in the Quillen spectral sequence for all rr. Furthermore, we know that the primitive element [ω5,ω9]=[ω5|ω9][ω9|ω5]E18,8Q[\omega^{5},\omega^{9}]=[\omega^{5}|\omega^{9}]-[\omega^{9}|\omega^{5}]\in{}^{Q}\!E_{1}^{8,8} of degree 1616, which is non-zero by Corollary 6.3, does not appear in the abutment. Since it is annihilated by all differentials drQ{}^{Q}d_{r}, it must be in the image of a class of degree 15 in genus 7\leq 7. The only such class is [ω5ω9][\omega^{5}\wedge\omega^{9}] in genus 55. We conclude that

drQ([ω5ω9])={α[ω5,ω9] if r=30 else{}^{Q}d_{r}\left([\omega^{5}\wedge\omega^{9}]\right)=\begin{cases}\alpha[\omega^{5},\omega^{9}]\qquad\hbox{ if }r=3\\ 0\qquad\qquad\qquad\hbox{ else}\end{cases}

for some α×\alpha\in\mathbb{Q}^{\times}. Since [ω5],[ω9][\omega^{5}],[\omega^{9}] generate a free Lie algebra in Prim(E1Q)\operatorname{Prim}\big{(}{}^{Q}\!E_{1}\big{)} (Corollary 6.3), we may deduce the non-vanishing of many more elements, and the non-triviality of Lie brackets involving the generator ω5ω9\omega^{5}\wedge\omega^{9}, which is noteworthy as ω5ω9\omega^{5}\wedge\omega^{9} has odd degree.

As a warm-up example, let us first show that [ω5,ω5ω9]E18,13Q[\omega^{5},\omega^{5}\wedge\omega^{9}]\in{}^{Q}\!E_{1}^{8,13} is nonzero. As argued above, both ω5\omega^{5} and ω5ω9\omega^{5}\wedge\omega^{9} represent nonzero classes on E3Q{}^{Q}\!E_{3}, and the Leibniz rule gives d3Q[ω5,ω5ω9]=[ω5,[ω5,ω9]]{}^{Q}d_{3}[\omega^{5},\omega^{5}\wedge\omega^{9}]=[\omega^{5},[\omega^{5},\omega^{9}]], which is nonzero by freeness of the Lie algebra generated by [ω5][\omega^{5}] and [ω9][\omega^{9}]. Therefore [ω5,ω5ω9]0[\omega^{5},\omega^{5}\wedge\omega^{9}]\neq 0. More generally, we can obtain an infinite supply of linearly independent primitive elements by writing down a Hall basis for the free Lie algebra on two generators 𝕃(x,y)\mathbb{L}(x,y), choosing for each one an occurence of [x,y][x,y] and replacing it with ω5ω9\omega^{5}\wedge\omega^{9}, and finally replacing xx with ω5\omega^{5} and yy with ω9\omega^{9}. For example:

(92) (xy)x,(xy)y\displaystyle(xy)x\,,\,(xy)y \displaystyle\rightarrow [ω5ω9,ω5],[ω5ω9,ω9]\displaystyle[\omega^{5}\wedge\omega^{9},\omega^{5}]\,,\,[\omega^{5}\wedge\omega^{9},\omega^{9}]
((xy)x)x,((xy)y)x),((xy)y)y\displaystyle((xy)x)x\,,\,((xy)y)x)\,,\,((xy)y)y \displaystyle\rightarrow [[ω5ω9,ω5],ω5],[[ω5ω9,ω9],ω5],[[ω5ω9,ω9],ω9]\displaystyle[[\omega^{5}\wedge\omega^{9},\omega^{5}],\omega^{5}]\,,\,[[\omega^{5}\wedge\omega^{9},\omega^{9}],\omega^{5}]\,,\,[[\omega^{5}\wedge\omega^{9},\omega^{9}],\omega^{9}]

where, in standard notation (ab)(ab) denotes [a,b][a,b]. In length four the Hall basis

((xy)x)x)x,((xy)y)x)x,((xy)y)y)x,((xy)y)y)y,((xy)y)(xy),((xy)x)(xy)((xy)x)x)x\ ,\ ((xy)y)x)x\ ,\ ((xy)y)y)x\ ,\ ((xy)y)y)y\ ,\ ((xy)y)(xy)\ ,\ ((xy)x)(xy)

may be lifted to, for example, a set of six elements:

[[[ω5ω9,ω5],ω5],ω5],[[[ω5ω9,ω9],ω5],ω5],[[[ω5ω9,ω9],ω9],ω5][[[\omega^{5}\wedge\omega^{9},\omega^{5}],\omega^{5}],\omega^{5}]\ ,\ [[[\omega^{5}\wedge\omega^{9},\omega^{9}],\omega^{5}],\omega^{5}]\ ,\ [[[\omega^{5}\wedge\omega^{9},\omega^{9}],\omega^{9}],\omega^{5}]
[[[ω5ω9,ω9],ω9],ω9],[[ω5,ω9],ω9],ω5ω9],[[ω5,ω9],ω5],ω5ω9][[[\omega^{5}\wedge\omega^{9},\omega^{9}],\omega^{9}],\omega^{9}]\ ,\ [[\omega^{5},\omega^{9}],\omega^{9}],\omega^{5}\wedge\omega^{9}]\ ,\ [[\omega^{5},\omega^{9}],\omega^{5}],\omega^{5}\wedge\omega^{9}]

They are independent since their images under d3Q{}^{Q}d_{3} are part of a Hall basis for the free lie algebra 𝕃(ω5,ω9)\mathbb{L}(\omega^{5},\omega^{9}), and hence are independent. In particular, since the Lie bracket [ω5ω9,ω5][\omega^{5}\wedge\omega^{9},\omega^{5}] is non-zero, the Milnor–Moore theorem implies that the two classes in genus 88 and degree 21 of the form [ω5|ω5ω9][\omega^{5}|\omega^{5}\wedge\omega^{9}], [ω5ω9|ω5][\omega^{5}\wedge\omega^{9}|\omega^{5}] are linearly independent, which proves Conjecture 1.16 up to and including genus 99.

In genus 1010, injectivity of the map T(Ωc[1])W0Hc(𝒜;)T(\Omega^{*}_{c}[-1])\to W_{0}H^{*}_{c}(\mathcal{A};\mathbb{R}) in Conjecture 1.16 is not known for four cohomological degrees, namely d=34,30,29,d=34,30,29, and 2525. Conjecture 1.16 predicts the existence of 2,1,2,2,1,2, and 44 independent classes respectively, while we prove that the dimension of W0Hcd(𝒜10;)W_{0}H^{d}_{c}(\mathcal{A}_{10};\mathbb{R}) is at least 1,0,1,1,0,1, and 33 in these degrees, respectively. For example, in degree 3030, it is not known if the primitive element [ω5ω9|ω5ω9][\omega^{5}\wedge\omega^{9}|\omega^{5}\wedge\omega^{9}] is non-zero. See Table 1.

7.7. Proof of Theorem 1.19

Following the notation suggested by Namikawa, we shall write 𝒜g\mathcal{A}_{g}^{\textrm{\begin{CJK}{UTF8}{min}サ\end{CJK}}} for the minimal Satake, or Baily–Borel, compactification of 𝒜g\mathcal{A}_{g}.

Proof.

The Satake compactification of 𝒜h\mathcal{A}_{h} has a natural stratification

𝒜h=gh𝒜g.\mathcal{A}_{h}^{\textrm{\begin{CJK}{UTF8}{min}サ\end{CJK}}}=\bigsqcup_{g\leq h}\mathcal{A}_{g}.

This induces a spectral sequence of mixed Hodge structures E{}^{\textrm{\begin{CJK}{UTF8}{min}サ\end{CJK}}}\!E_{*} abutting to the cohomology H(𝒜h)H^{*}(\mathcal{A}_{h}^{\textrm{\begin{CJK}{UTF8}{min}サ\end{CJK}}}) whose E1E_{1}-page is

E1g,k=Hcg+k(𝒜g),{}^{\textrm{\begin{CJK}{UTF8}{min}サ\end{CJK}}}\!E_{1}^{g,k}=H^{g+k}_{c}(\mathcal{A}_{g}),

for ghg\leq h, and 0 otherwise. Let us consider the E1E_{1}-page of the induced spectral sequence on the weight zero subspaces. Choose h=6h=6, which is sufficiently large so that H6(𝒜h)H^{6}(\mathcal{A}_{h}^{\textrm{\begin{CJK}{UTF8}{min}サ\end{CJK}}}) is stable. Then E1{}^{\textrm{\begin{CJK}{UTF8}{min}サ\end{CJK}}}\!E_{1} agrees with the truncation of W0Hc(𝒜)W_{0}H^{*}_{c}(\mathcal{A}) after the g=6g=6 column, which is depicted in Table 1. Similarly, the corresponding spectral sequence abutting to W0H(𝒜3)W_{0}H^{*}(\mathcal{A}_{3}^{\textrm{\begin{CJK}{UTF8}{min}サ\end{CJK}}}) is obtained by truncating after the third column. For degree reasons, both of these groups are given by E13,3{}^{\textrm{\begin{CJK}{UTF8}{min}サ\end{CJK}}}\!E_{1}^{3,3}. Thus, we have isomorphisms

W0H6(𝒜6)W0Hc6(𝒜3)W0H6(𝒜3).W_{0}H^{6}(\mathcal{A}_{6}^{\textrm{\begin{CJK}{UTF8}{min}サ\end{CJK}}})\cong W_{0}H^{6}_{c}(\mathcal{A}_{3})\cong W_{0}H^{6}(\mathcal{A}_{3}^{\textrm{\begin{CJK}{UTF8}{min}サ\end{CJK}}}).

In particular, the inclusion 𝒜3𝒜6\mathcal{A}_{3}^{\textrm{\begin{CJK}{UTF8}{min}サ\end{CJK}}}\subset\mathcal{A}_{6}^{\textrm{\begin{CJK}{UTF8}{min}サ\end{CJK}}} induces an isomorphism on W0H6W_{0}H^{6}. Now we show that the induced map on all of H6H^{6} is injective. To see this, note that H6(𝒜6)H^{6}(\mathcal{A}_{6}^{\textrm{\begin{CJK}{UTF8}{min}サ\end{CJK}}}) has rank 22, with

H6(𝒜6)/W0H6(𝒜6)H^{6}(\mathcal{A}_{6}^{\textrm{\begin{CJK}{UTF8}{min}サ\end{CJK}}})/W_{0}H^{6}(\mathcal{A}_{6}^{\textrm{\begin{CJK}{UTF8}{min}サ\end{CJK}}})

spanned by the Goresky–Pardon lift c~3\tilde{c}_{3} of c3(Λ)c_{3}(\Lambda), the third Chern class of the Hodge bundle [CL17, Loo17]. Thus, it will suffice to show that the restriction of c~3\tilde{c}_{3} to 𝒜3\mathcal{A}_{3}^{\textrm{\begin{CJK}{UTF8}{min}サ\end{CJK}}} is not zero.

The Goresky–Pardon lift c~3\tilde{c}_{3} has the following property: fix gg and choose a toroidal compactification 𝒜gΣ\mathcal{A}_{g}^{\Sigma}, as in [AMRT75]. Let Λg\Lambda_{g} denote the Hodge bundle on 𝒜g\mathcal{A}_{g} and let Λ~g\widetilde{\Lambda}_{g} be the extension to 𝒜gΣ\mathcal{A}_{g}^{\Sigma} constructed in [Mum77]. Then there is a unique projection πΣ:𝒜gΣ𝒜g\pi_{\Sigma}\colon\mathcal{A}_{g}^{\Sigma}\to\mathcal{A}_{g}^{\textrm{\begin{CJK}{UTF8}{min}サ\end{CJK}}} that extends the identity on the shared open subset 𝒜g\mathcal{A}_{g}, and

(93) πΣ(c~3)=c3(Λ~g).\pi_{\Sigma}^{*}(\tilde{c}_{3})=c_{3}(\widetilde{\Lambda}_{g}).

See [GP02]. We can choose a toroidal compactification so that the Torelli map extends to a morphism from the moduli space of stable curves ¯g𝒜gΣ\overline{\mathcal{M}}_{g}\to\mathcal{A}_{g}^{\Sigma} and the pullback of c~3\tilde{c}_{3} is the Hodge class λ3\lambda_{3}. Then λ30\lambda_{3}\neq 0 in H6(¯3)H^{6}(\overline{\mathcal{M}}_{3}) because it appears as a multiplicative factor in the integrand of explicit nonzero Hodge integrals [FP00]. It follows that c~3\tilde{c}_{3} is nonzero in H6(𝒜3),H^{6}(\mathcal{A}_{3}^{\textrm{\begin{CJK}{UTF8}{min}サ\end{CJK}}}), which proves the claim.

Next, we claim that c~3\tilde{c}_{3} is in the image of the push-forward for the open inclusion

ι:Hc6(𝒜3)H6(𝒜3).\iota_{*}\colon H^{6}_{c}(\mathcal{A}_{3})\to H^{6}(\mathcal{A}_{3}^{\textrm{\begin{CJK}{UTF8}{min}サ\end{CJK}}}).

The property (93) for g=2g=2 shows that the restriction of c~3\tilde{c}_{3} to 𝒜2\mathcal{A}_{2}^{\textrm{\begin{CJK}{UTF8}{min}サ\end{CJK}}} vanishes, since Λ2\Lambda_{2} has rank 2. Applying excision gives a long exact sequence

Hc6(𝒜3)H6(𝒜3)H6(𝒜2)\cdots\to H^{6}_{c}(\mathcal{A}_{3})\to H^{6}(\mathcal{A}^{\textrm{\begin{CJK}{UTF8}{min}サ\end{CJK}}}_{3})\to H^{6}(\mathcal{A}^{\textrm{\begin{CJK}{UTF8}{min}サ\end{CJK}}}_{2})\to\cdots

Since c~3\tilde{c}_{3} is in the kernel of the map to H6(𝒜2)H^{6}(\mathcal{A}_{2}^{\textrm{\begin{CJK}{UTF8}{min}サ\end{CJK}}}), it must be in the image of Hc6(𝒜3)H^{6}_{c}(\mathcal{A}_{3}), as claimed.

We have shown that the rank two stable cohomology group H6(𝒜6)H^{6}(\mathcal{A}_{6}^{\textrm{\begin{CJK}{UTF8}{min}サ\end{CJK}}}) injects into H6(𝒜3)H^{6}(\mathcal{A}_{3}^{\textrm{\begin{CJK}{UTF8}{min}サ\end{CJK}}}) and the image is contained in the image of Hc6(𝒜3)H^{6}_{c}(\mathcal{A}_{3}), which also has rank 2 [Hai02]. Thus we have a zig-zag of isomorphisms of mixed Hodge structures

Hc6(𝒜3)Im(H6(𝒜6)H6(𝒜3))H6(𝒜)H^{6}_{c}(\mathcal{A}_{3})\xrightarrow{\sim}\ \mathrm{Im}\,\big{(}H^{6}(\mathcal{A}_{6}^{\textrm{\begin{CJK}{UTF8}{min}サ\end{CJK}}})\to H^{6}(\mathcal{A}_{3}^{\textrm{\begin{CJK}{UTF8}{min}サ\end{CJK}}})\big{)}\ \xleftarrow{\sim}H^{6}(\mathcal{A}_{\infty}^{\textrm{\begin{CJK}{UTF8}{min}サ\end{CJK}}})

Since they come from algebraic maps, they also induce isomorphisms in algebraic de Rham cohomology, and hence remain true in any suitable category of realisations. The main result of [Loo17] shows that H6(𝒜6)H^{6}(\mathcal{A}_{6}^{\textrm{\begin{CJK}{UTF8}{min}サ\end{CJK}}}) is a nontrivial extension of (3)\mathbb{Q}(-3) by \mathbb{Q}, whose extension class is given by a nonzero rational multiple of ζ(3)\zeta(3). This means that, in a suitable choice of basis for Betti and de Rham cohomology, the period matrix PP is upper-triangular with ζ(3)\zeta(3) above the diagonal, and 11, (2πi)3(2\pi i)^{3} along the diagonal. Following the zig-zag and applying Poincaré duality to Hc6(𝒜3)H^{6}_{c}(\mathcal{A}_{3}) shows that H6(𝒜3)H^{6}(\mathcal{A}_{3}) is likewise a nontrivial extension of (6)\mathbb{Q}(-6) by (3)\mathbb{Q}(-3). More precisely, its period matrix, with respect to the dual basis, is the inverse transpose of PP times (2πi)6(2\pi i)^{6}, and hence lower-triangular with ζ(3)(2πi)3-\zeta(3)(2\pi i)^{3} below the diagonal. ∎

References

  • [ACP22] Daniel Allcock, Daniel Corey, and Sam Payne, Tropical moduli spaces as symmetric Δ\Delta-complexes, Bull. Lond. Math. Soc. 54 (2022), no. 1, 193–205. MR 4396931
  • [AMP24] Avner Ash, Jeremy Miller, and Peter Patzt, Hopf algebras, Steinberg modules, and the unstable cohomology of SLn()\mathrm{SL}_{n}(\mathbb{Z}) and GLn()\mathrm{GL}_{n}(\mathbb{Z}), arXiv:2404.13776, 2024.
  • [AMRT75] Avner Ash, David Mumford, Michael Rapoport, and Yung-sheng Tai, Smooth compactification of locally symmetric varieties, Math. Sci. Press, Brookline, Mass., 1975, Lie Groups: History, Frontiers and Applications, Vol. IV. MR 0457437
  • [Ash94] Avner Ash, Unstable cohomology of SL(n,𝒪){\rm SL}(n,\mathcal{O}), J. Algebra 167 (1994), no. 2, 330–342. MR 1283290
  • [BBC+24] Madeline Brandt, Juliette Bruce, Melody Chan, Margarida Melo, Gwyneth Moreland, and Corey Wolfe, On the top-weight rational cohomology of 𝒜g\mathcal{A}_{g}, Geom. Topol. 28 (2024), no. 2, 497–538. MR 4718121
  • [BBV10] Johannes Blümlein, David J. Broadhurst, and Jos A. M. Vermaseren, The multiple zeta value data mine, Computer Physics Communications 181 (2010), no. 3, 582–625.
  • [BCG23] George Boxer, Frank Calegari, and Toby Gee, Cuspidal cohomology classes for GLn()\mathrm{GL}_{n}(\mathbb{Z}), 2023, arXiv:2309.15944.
  • [BHM93] Marcel Bökstedt, Wu-Chung Hsiang, and Ib Madsen, The cyclotomic trace and algebraic KK-theory of spaces, Invent. Math. 111 (1993), no. 3, 465–539. MR 1202133
  • [BMP+24] Benjamin Brück, Jeremy Miller, Peter Patzt, Robin J. Sroka, and Jennifer C. H. Wilson, On the codimension-two cohomology of SLn(){\rm SL}_{n}(\mathbb{Z}), Adv. Math. 451 (2024), Paper No. 109795. MR 4764368
  • [BMV11] Silvia Brannetti, Margarida Melo, and Filippo Viviani, On the tropical Torelli map, Adv. Math. 226 (2011), no. 3, 2546–2586. MR 2739784
  • [BPS23] Benjamin Brück, Peter Patzt, and Robin J. Sroka, A presentation of symplectic Steinberg modules and cohomology of Sp2n()\mathrm{Sp}_{2n}(\mathbb{Z}), arXiv:2306.03180, 2023.
  • [Bro12] Francis Brown, Mixed Tate motives over \mathbb{Z}, Ann. of Math. (2) 175 (2012), no. 2, 949–976. MR 2993755
  • [Bro21a] by same author, Depth-graded motivic multiple zeta values, Compos. Math. 157 (2021), no. 3, 529–572. MR 4236194
  • [Bro21b] by same author, Invariant differential forms on complexes of graphs and Feynman integrals, SIGMA Symmetry Integrability Geom. Methods Appl. 17 (2021), Paper No. 103, 54. MR 4343257
  • [Bro23] by same author, Bordifications of the moduli spaces of tropical curves and abelian varieties, and unstable cohomology of GLg()\mathrm{GL}_{g}(\mathbb{Z}) and SLg()\mathrm{SL}_{g}(\mathbb{Z}), arXiv:2309.12753, 2023.
  • [BS73] Armand Borel and Jean-Pierre Serre, Corners and arithmetic groups, Comment. Math. Helv. 48 (1973), 436–491. MR 387495
  • [BS24] Francis Brown and Oliver Schnetz, The wheel classes in the locally finite homology of GLn()\mathrm{GL}_{n}(\mathbb{Z}), canonical integrals and zeta values, 2024.
  • [CE99] Henri Cartan and Samuel Eilenberg, Homological algebra, Princeton Landmarks in Mathematics, Princeton University Press, Princeton, NJ, 1999, With an appendix by David A. Buchsbaum, Reprint of the 1956 original. MR 1731415
  • [CFP12] Thomas Church, Benson Farb, and Andrew Putman, The rational cohomology of the mapping class group vanishes in its virtual cohomological dimension, Int. Math. Res. Not. IMRN (2012), no. 21, 5025–5030. MR 2993444
  • [CFP14] by same author, A stability conjecture for the unstable cohomology of SLn{\rm SL}_{n}\mathbb{Z}, mapping class groups, and Aut(Fn){\rm Aut}(F_{n}), Algebraic topology: applications and new directions, Contemp. Math., vol. 620, Amer. Math. Soc., Providence, RI, 2014, pp. 55–70. MR 3290086
  • [CFvdG20] Fabien Cléry, Carel Faber, and Gerard van der Geer, Concomitants of ternary quartics and vector-valued Siegel and Teichmüller modular forms of genus three, Selecta Math. (N.S.) 26 (2020), no. 4, Paper No. 55, 39. MR 4125987
  • [CGP21] Melody Chan, Søren Galatius, and Sam Payne, Tropical curves, graph complexes, and top weight cohomology of g\mathcal{M}_{g}, J. Amer. Math. Soc. 34 (2021), 565–594.
  • [CHKV16] James Conant, Allen Hatcher, Martin Kassabov, and Karen Vogtmann, Assembling homology classes in automorphism groups of free groups, Comment. Math. Helv. 91 (2016), no. 4, 751–806. MR 3566523
  • [CK98] Alain Connes and Dirk Kreimer, Hopf algebras, renormalization and noncommutative geometry, Comm. Math. Phys. 199 (1998), no. 1, 203–242. MR 1660199
  • [CL17] Jiaming Chen and Eduard Looijenga, The stable cohomology of the Satake compactification of 𝒜g\mathcal{A}_{g}, Geom. Topol. 21 (2017), no. 4, 2231–2241. MR 3654107
  • [CP17] Thomas Church and Andrew Putman, The codimension-one cohomology of SLn{\rm SL}_{n}\mathbb{Z}, Geom. Topol. 21 (2017), no. 2, 999–1032. MR 3626596
  • [CV86] Marc Culler and Karen Vogtmann, Moduli of graphs and automorphisms of free groups, Invent. Math. 84 (1986), no. 1, 91–119. MR 830040
  • [Del89] P. Deligne, Le groupe fondamental de la droite projective moins trois points, Galois groups over 𝐐{\bf Q} (Berkeley, CA, 1987), Math. Sci. Res. Inst. Publ., vol. 16, Springer, New York, 1989, pp. 79–297. MR 1012168
  • [Dou59a] Adrien Douady, La suite spectrale d’Adams, Séminaire Henri Cartan 11 (1958-1959), no. 2, 1–18.
  • [Dou59b] by same author, La suite spectrale d’Adams: structure multiplicative, Séminaire Henri Cartan 11 (1958-1959), no. 2, 1–13, talk:19.
  • [DSEVKM19] Mathieu Dutour Sikirić, Philippe Elbaz-Vincent, Alexander Kupers, and Jacques Martinet, Voronoi complexes in higher dimensions, cohomology of GLN(Z){GL}_{N}({Z}) for N8{N}\geq 8 and the triviality of K8(Z){K}_{8}({Z}), 2019.
  • [EVGS13] Philippe Elbaz-Vincent, Herbert Gangl, and Christophe Soulé, Perfect forms, K-theory and the cohomology of modular groups, Adv. Math. 245 (2013), 587–624. MR 3084439
  • [FC90] Gerd Faltings and Ching-Li Chai, Degeneration of abelian varieties, Ergebnisse der Mathematik und ihrer Grenzgebiete (3) [Results in Mathematics and Related Areas (3)], vol. 22, Springer-Verlag, Berlin, 1990, With an appendix by David Mumford. MR 1083353
  • [FP00] Carel Faber and Rahul Pandharipande, Hodge integrals and Gromov-Witten theory, Invent. Math. 139 (2000), no. 1, 173–199. MR 1728879
  • [FRSS18] Tyler Foster, Joseph Rabinoff, Farbod Shokrieh, and Alejandro Soto, Non-Archimedean and tropical theta functions, Math. Ann. 372 (2018), no. 3-4, 891–914. MR 3880286
  • [GCKT20a] Imma Gálvez-Carrillo, Ralph M. Kaufmann, and Andrew Tonks, Three Hopf algebras from number theory, physics & topology, and their common background I: operadic & simplicial aspects, Commun. Number Theory Phys. 14 (2020), no. 1, 1–90. MR 4060970
  • [GCKT20b] by same author, Three Hopf algebras from number theory, physics & topology, and their common background II: general categorical formulation, Commun. Number Theory Phys. 14 (2020), no. 1, 91–169. MR 4060971
  • [GHT18] Samuel Grushevsky, Klaus Hulek, and Orsola Tommasi, Stable Betti numbers of (partial) toroidal compactifications of the moduli space of abelian varieties, Geometry and physics. Vol. II, Oxford Univ. Press, Oxford, 2018, With an appendix by Mathieu Dutour Sikirić, pp. 581–609. MR 3931788
  • [GKRW20] Soren Galatius, Alexander Kupers, and Oscar Randal-Williams, EE_{\infty}-cells and general linear groups of infinite fields, arXiv:2005.05620, 2020.
  • [Goe] Sebastian Goette (https://mathoverflow.net/users/70808/sebastian-goette), Multiplicative structure on spectral sequence, MathOverflow, URL:https://mathoverflow.net/q/231235 (version: 2016-02-15).
  • [GP02] Mark Goresky and William Pardon, Chern classes of automorphic vector bundles, Invent. Math. 147 (2002), no. 3, 561–612. MR 1893006
  • [Gro13] Harald Grobner, Residues of Eisenstein series and the automorphic cohomology of reductive groups, Compos. Math. 149 (2013), no. 7, 1061–1090. MR 3078638
  • [Gru09] Samuel Grushevsky, Geometry of 𝒜g\mathcal{A}_{g} and its compactifications, Algebraic geometry—Seattle 2005. Part 1, Proc. Sympos. Pure Math., vol. 80, Amer. Math. Soc., Providence, RI, 2009, pp. 193–234. MR 2483936
  • [Gun00] Paul E. Gunnells, Symplectic modular symbols, Duke Math. J. 102 (2000), no. 2, 329–350. MR 1749441
  • [Hai02] Richard Hain, The rational cohomology ring of the moduli space of abelian 3-folds, Math. Res. Lett. 9 (2002), no. 4, 473–491. MR 1928867
  • [Har86] John L. Harer, The virtual cohomological dimension of the mapping class group of an orientable surface, Invent. Math. 84 (1986), no. 1, 157–176. MR 830043
  • [Hat95] Allen Hatcher, Homological stability for automorphism groups of free groups, Comment. Math. Helv. 70 (1995), no. 1, 39–62. MR 1314940
  • [Hat02] by same author, Algebraic topology, Cambridge University Press, Cambridge, 2002. MR 1867354
  • [Hor05] Ivan Horozov, Euler characteristics of arithmetic groups, Math. Res. Lett. 12 (2005), no. 2-3, 275–291. MR 2150884
  • [HT12] Klaus Hulek and Orsola Tommasi, Cohomology of the second Voronoi compactification of 𝒜4\mathcal{A}_{4}, Doc. Math. 17 (2012), 195–244. MR 2946823
  • [HT18] by same author, The topology of 𝒜g\mathcal{A}_{g} and its compactifications, Geometry of moduli, Abel Symp., vol. 14, Springer, Cham, 2018, With an appendix by Olivier Taïbi, pp. 135–193. MR 3968046
  • [HV98] Allen Hatcher and Karen Vogtmann, Rational homology of Aut(Fn){\rm Aut}(F_{n}), Math. Res. Lett. 5 (1998), no. 6, 759–780. MR 1671188
  • [HV04] by same author, Homology stability for outer automorphism groups of free groups, Algebr. Geom. Topol. 4 (2004), 1253–1272. MR 2113904
  • [Jan24] Mikala Jansen, Unstable algebraic K-theory: homological stability and other observations, preprint, arXiv:2405.02065, 2024.
  • [KW17] Ralph M. Kaufmann and Benjamin C. Ward, Feynman categories, Astérisque (2017), no. 387, vii+161. MR 3636409
  • [KW23] by same author, Koszul Feynman categories, Proc. Amer. Math. Soc. 151 (2023), no. 8, 3253–3267. MR 4591764
  • [KWŽ17] Anton Khoroshkin, Thomas Willwacher, and Marko Živković, Differentials on graph complexes, Adv. Math. 307 (2017), 1184–1214. MR 3590540
  • [Lee78] Ronnie Lee, On unstable cohomology classes of SLn()\mathrm{SL}_{n}(\mathbb{Z}), Proc. Nat. Acad. Sci. U.S.A. 75 (1978), no. 1, 43–44. MR 467778
  • [Loo17] Eduard Looijenga, Goresky-Pardon lifts of Chern classes and associated Tate extensions, Compos. Math. 153 (2017), no. 7, 1349–1371. MR 3705260
  • [LS76] Ronnie Lee and Robert H. Szczarba, On the homology and cohomology of congruence subgroups, Invent. Math. 33 (1976), no. 1, 15–53. MR 422498
  • [LS19] Jian-Shu Li and Binyong Sun, Low degree cohomologies of congruence groups, Sci. China Math. 62 (2019), no. 11, 2287–2308. MR 4028274
  • [LV12] Jean-Louis Loday and Bruno Vallette, Algebraic operads, Grundlehren der mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], vol. 346, Springer, Heidelberg, 2012. MR 2954392
  • [Mil12] James S. Milne, Basic theory of affine group schemes, 2012, Available at www.jmilne.org/math/.
  • [MM65] John W. Milnor and John C. Moore, On the structure of Hopf algebras, Ann. of Math. (2) 81 (1965), 211–264. MR 174052
  • [MSS13] Shigeyuki Morita, Takuya Sakasai, and Masaaki Suzuki, Abelianizations of derivation Lie algebras of the free associative algebra and the free Lie algebra, Duke Math. J. 162 (2013), no. 5, 965–1002. MR 3047471
  • [Mum77] David Mumford, Hirzebruch’s proportionality theorem in the noncompact case, Invent. Math. 42 (1977), 239–272. MR 471627
  • [Nam80] Yukihiko Namikawa, Toroidal compactification of Siegel spaces, Lecture Notes in Mathematics, vol. 812, Springer, Berlin, 1980. MR 584625
  • [OO21] Yuji Odaka and Yoshiki Oshima, Collapsing K3 surfaces, tropical geometry and moduli compactifications of Satake, Morgan-Shalen type, MSJ Memoirs, vol. 40, Mathematical Society of Japan, Tokyo, 2021. MR 4290084
  • [PW21] Sam Payne and Thomas Willwacher, Weight two compactly supported cohomology of moduli spaces of curves, arXiv:2110.05711. To appear in Duke Math. J., 2021.
  • [PW24] by same author, Weight 11 Compactly Supported Cohomology of Moduli Spaces of Curves, Int. Math. Res. Not. IMRN (2024), no. 8, 7060–7098. MR 4735654
  • [Qui69] Daniel Quillen, Rational homotopy theory, Ann. of Math. (2) 90 (1969), 205–295. MR 258031
  • [Qui73] by same author, Higher algebraic KK-theory. I, Algebraic KK-theory, I: Higher KK-theories (Proc. Conf., Battelle Memorial Inst., Seattle, Wash., 1972), Lecture Notes in Math., vol. Vol. 341, Springer, Berlin-New York, 1973, pp. 85–147. MR 338129
  • [Rog] John Rognes, Spectral sequences, https://www.uio.no/studier/emner/matnat/math/MAT9580/v21/dokumenter/spseq.pdf.
  • [RW14] Carlo A. Rossi and Thomas Willwacher, P. Etingof’s conjecture about Drinfeld associators, arXiv:1404.2047, 2014.
  • [Seg73] Graeme Segal, Configuration-spaces and iterated loop-spaces, Invent. Math. 21 (1973), 213–221.
  • [vdG99] Gerard van der Geer, Cycles on the moduli space of abelian varieties, Moduli of curves and abelian varieties, Aspects Math., E33, Friedr. Vieweg, Braunschweig, 1999, pp. 65–89. MR 1722539
  • [vdGL21] Gerard van der Geer and Eduard Looijenga, Lifting Chern classes by means of Ekedahl-Oort strata, Tunis. J. Math. 3 (2021), no. 3, 469–480. MR 4258631
  • [Wal85] Friedhelm Waldhausen, Algebraic K-theory of spaces, Algebraic and Geometric Topology (Berlin, Heidelberg) (Andrew Ranicki, Norman Levitt, and Frank Quinn, eds.), Springer Berlin Heidelberg, 1985, pp. 318–419.
  • [Wei94] Charles A. Weibel, An introduction to homological algebra, Cambridge Studies in Advanced Mathematics, vol. 38, Cambridge University Press, Cambridge, 1994. MR 1269324
  • [Wei13] by same author, The KK-book, Graduate Studies in Mathematics, vol. 145, American Mathematical Society, Providence, RI, 2013, An introduction to algebraic KK-theory. MR 3076731
  • [Wil15] Thomas Willwacher, M. Kontsevich’s graph complex and the Grothendieck-Teichmüller Lie algebra, Invent. Math. 200 (2015), no. 3, 671–760. MR 3348138
  • [Yua23] Allen Yuan, Integral models for spaces via the higher Frobenius, J. Amer. Math. Soc. 36 (2023), no. 1, 107–175.
  • [Yun22] Claudia He Yun, Discrete Morse theory for symmetric Delta-complexes, arXiv:2209.01070, 2022.
  • [Zag93] Don Zagier, Periods of modular forms, traces of Hecke operators, and multiple zeta values, Research into automorphic forms and LL functions (Kyoto, 1992), no. 843, Kyoto Univ., 1993, pp. 162–170. MR 1296720