Cohomology of automorphism groups of free groups with twisted coefficients
Abstract.
We compute the groups and in a stable range, where is obtained by applying a Schur functor to or , respectively the first rational homology and cohomology of . The answer may be described in terms of stable multiplicities of irreducibles in the plethysm of symmetric powers.
We also compute the stable integral cohomology groups of with coefficients in or .
Key words and phrases:
Automorphisms of free groups, homology stability2010 Mathematics Subject Classification:
20F28, 20J06, 57R201. Statement of results
Galatius [12] has proved the remarkable theorem that the natural homomorphisms
both induce homology isomorphisms in degrees with integral coefficients. His approach is to model as the space of graphs of the homotopy type of , and as the space of pointed graphs of the same homotopy type. He then produces a natural map from such spaces of graphs to the infinite loop space , which he shows has a certain homological connectivity. One consequence of this is that for .
At the same time, Satoh [26, 27] has studied the low dimensional (co)homology of and with coefficients in the module given by the abelianisation of , and in the dual module . His methods are those of combinatorial group theory, and proceed by calculation with a presentation of these groups.
Our goal is to show that the stable cohomology of and with twisted coefficients may also be approached with the geometric techniques used by Galatius, along with a little representation theory. The method we shall introduce is quite general and may be applied whenever a suitable Madsen–Weiss-type theorem has been proved; in Appendix B we will show how to use it to recover a theorem of Looijenga [20] on the stable cohomology of mapping class groups with twisted coefficients.
For and we will consider cohomology with coefficients in the modules and , and more generally with coefficients in and , where is the Schur functor associated to a partition , which we think of as being given by a Young diagram. To define this, recall that to such a partition there is an associated irreducible -module , the Specht module. For a -vector space we may consider as a -module by permuting the factors, and we may hence form . This construction defines the Schur functor . It is a basic result that is an irreducible representation of .
In fact, our result is best expressed as calculating as a -module. The result for may then be extracted as
We write for the sign representation of .
Theorem A.
-
(i)
for .
-
(ii)
for if , and is the permutation module on the set of partitions of .
Theorem A (i) may also be deduced from work of Djament–Vespa [9]. We believe that Theorem A (ii) may also be obtained by combining work of Djament [8] and Vespa [28].
However, more important than these particular results is our technique, which is of very general applicability. For example, it can easily be modified to obtain results for .
Theorem B.
-
(i)
for .
-
(ii)
for if , and as long as is the permutation module on the set of partitions of having no parts of size 1.
We give tables listing the dimensions of the groups and for in Appendix C.
Each partition of may be expressed as a partition of a smaller set with no parts of size 1 along with the set of parts of size 1, which translates to the expression
as long as . Applying and using the Pieri rule gives the pleasant formula
(1.1) |
as long as , where denotes the set of Young diagrams which may be obtained from by removing at most one box from each row.
1.1. Stable plethysm of symmetric powers
For a Young diagram and integers and such that , let be the Young diagram obtained by adding a row of length to the top of , and let denote the multiplicity of the irreducible -representation in . Manivel has shown [21] that the numbers are increasing and eventually constant functions of both and , and we write for the stable value. This stable value is attained as soon as and . Using the work of Manivel, we are able to relate these stable multiplicities directly to the cohomology of , as follows.
Theorem C.
As long as we have
and as long as we have
It follows directly from [21, Proposition 4.4.1] that if , or if and does not consist of two rows of equal length.
1.2. Symmetric and exterior powers
In order to demonstrate how our theorems may be used, we compute the dimension of the associated cohomology groups for the modules and .
Corollary D.
-
(i)
For , has dimension given by the number of partitions of ; for , has dimension given by the number of partitions of into pieces none of which are 1.
-
(ii)
Let . For ,
Proof.
The dimension of is the multiplicity of the sign representation in , which by Theorem A is the multiplicity of the trivial representation in the permutation module for the set of partitions of : this is the number of partitions of .
The dimension of is the multiplicity of the trivial representation in , which by Theorem A is the multiplicity of the sign representation in the permutation module for the set of partitions of . This may be computed via inner product of characters as times
but, writing for the stabiliser of a partition , we have that is times the multiplicity of the sign representation in the trivial representation of , which is zero (as always contains at least one transposition if ).
The arguments for are identical. ∎
1.3. Integral and torsion results
Finally, our technique can be made to give integral and local information as well. It is not hard to show that for (see Proposition 2.3), but we also have the following.
Theorem E.
-
(i)
is a free -module (on a single generator in degree 1).
-
(ii)
For a partition and a prime number , is a free -module (on generators in degree , the number of which may be deduced from Theorem A).
Acknowledgements
This paper is an update to a 2010 preprint [23] in which I proved, inter alia, Corollary D subject to a sequence of conjectures. Since that time, in joint work with Nathalie Wahl [25] we proved Conjecture A (and I explain in this paper how a version of Conjecture B follows from it), and in 2014 Søren Galatius explained to me a proof of Conjecture C. Thus these conjectural calculations from my 2010 preprint hold.
Recent work of Aurélien Djament [8] and Christine Vespa [28] obtains these calculations by very different means. I was motivated by their results to revisit these techniques and to clarify the status of the conjectures from my 2010 preprint. I would like to thank all of the above named for their interest in, and useful comments on, the content of this note.
2. An observation regarding homology stability
The observations of this section are no doubt known to some experts. The groups and fit into a more general family of groups denoted by Hatcher–Vogtmann [16], where and . Classifying spaces for these may be taken to be the spaces of graphs of the homotopy type of equipped with distinct ordered marked points.
Hatcher and Vogtmann [16] prove that the map (defined for ) that adds a loop at a particular marked point induces an integral homology isomorphism in degrees (and induces a rational homology isomorphism in degrees ). Furthermore, the map that forgets a marked point is an integral homology isomorphism in degrees (or if it is the last marked point).
We can make an immediate observation regarding cohomology with coefficients in from this homology stability result. There is an extension , and the corresponding Leray–Hochschild–Serre spectral sequence has two rows. However, as the projection is a homology equivalence in a range of degrees we deduce
Proposition 2.1.
The groups are zero for .
Similarly, there is a fibration with section and the projection map is a homology equivalence in a range of degrees, so
Proposition 2.2.
The groups are zero for .
More generally, the map has fibre and acts on its homology diagonally. Thus for the Serre spectral sequence has three rows, with
and
Using the fact that the projection map is a homology equivalence in degrees , we deduce that for . Continuing in this way for higher , we establish the following proposition.
Proposition 2.3.
For all , is zero for .
3. Homology stability with coefficient systems
By Galatius’ theorem, the groups are closely related to the symmetric groups, but also share many properties with mapping class groups of surfaces. These three families of groups are known to exhibit homological stability for integral homology, but symmetric groups and mapping class groups also exhibit homological stability for certain systems of coefficients, those of “finite degree”, a notion that is originally due to Dwyer [10] in his study of homological stability for general linear groups with coefficient systems.
This notion of degree may be formalised, in the context of free groups, as follows.
Definition 3.1.
Let denote the category whose objects are the finitely generated free groups, and where a morphism from to is given by a pair
consisting of an injective group homomorphism and a finitely-generated (free) subgroup such that .
A coefficient system is a covariant functor to the category of abelian groups. We declare the constant functors to be polynomial of degree 0, and more generally we declare a functor to be polynomial of degree if
-
(i)
sends the the canonical morphism to an injection, and
-
(ii)
the new coefficient system is polynomial of degree .
In a 2010 preprint [23], we conjectured (based on the analogy with general linear groups [10] and mapping class groups [17]) that the groups should exhibit homological stability in degrees when is a polynomial coefficient system of degree . Since then, in joint work with Wahl [25] we have established a quite general homological stability theorem with polynomial coefficients, and using the highly-connected simplicial complexes of [15] it applies in this case. The result obtained is as follows.
Theorem 3.2 (Randal-Williams–Wahl [25]).
If is a polynomial coefficient system of degree , then the natural map
induces an epimorphism for and an isomorphism for .
One advantage of the category over the more naïve category of finitely generated free groups and injective homomorphisms is that there are more functors out of it.
Definition 3.3.
-
(i)
Let be the coefficient system sending to , and sending a morphism to . There is an exact sequence
so this coefficient system is polynomial of degree 1.
-
(ii)
Let be the coefficient system sending to , and sending a morphism to the linear dual of
(Note that this does not define a functor on !) There is an exact sequence
so this coefficient system is polynomial of degree 1.
Remark 3.4.
Not only can more coefficient systems be defined on than on , but it follows from recent work of Djament–Vespa [9] that the most homologically interesting functors must be defined here: they show that if is a polynomial coefficient system which is reduced (i.e. ) and factors through , then
The coefficient system is reduced, but does not factor through : indeed, Satoh [26] has shown that , and we will show how to recover (the dual version of) this in Section 6.
There are some easy consequences of the definition of polynomiality which allow us to compute or estimate degrees of coefficient systems. Firstly, a summand of a polynomial functor of degree is again polynomial of degree , and an extension of two polynomial functors of degree is again polynomial of degree . Secondly, if and are coefficient systems which are polynomial of degree and respectively, and if they both take values in flat -modules, then is polynomial of degree . More generally, if is a field and are coefficient systems which are polynomial of degree and respectively and factor through the category of -modules, then is polynomial of degree .
We will often be interested in cohomology rather than homology, for which we will use the following result of Universal Coefficient-type, which is surely standard but for which we could not find a reference.
Lemma 3.5.
Let be a group, be a PID, and a left -module, and write , a right -module. There is a natural short exact sequence
Proof.
Analogous to [3, Proposition 7.1]. Let be a projective right -module resolution, so is a cochain complex of -modules which computes . Writing
and applying the Universal Coefficient Theorem (e.g. [4, p. 114]), using the fact that a submodule of a free module over a PID is free, gives the desired sequence. ∎
Corollary 3.6.
For all , for .
This proves Theorem A (i). We remark that Theorem A (i) also follows from [9, Théorème 1] and Lemma 3.5, because the functor is polynomial and vanishes on the trivial group (for ), so by that theorem vanishes in the stable range.
Finally, we record the stability range for cohomology with coefficients in with . This follows from , that is polynomial of degree as it is a summand of , the stability theorem, and Lemma 3.5.
Corollary 3.7.
The groups are independent of for .
4. Graphs labeled by a space
Let be a based space. Let us write for the topological category whose objects are finite sets inside , and whose morphisms from to consist of a real number and a graph with leaves and , equipped with a continuous map . This may be given a topology following [12]. Choosing once and for all an embedding the space may be defined as the subspace of consisting of those graphs which are connected and homotopy equivalent to .
Let be the subspace given by tuples where . Forgetting the point gives a map
which is a fibration with fibre over the point given by the graph . Furthermore, sending to defines a map
As the map is a fibration with homotopy-finite fibres, it admits a Becker–Gottlieb transfer map . Composing this with and taking the adjoint gives a map
to the free infinite loop space on , landing in the path component indexed by , the Euler characteristic of . This map is natural in , and factors through the map up to homotopy.
Theorem 4.1.
The induced map
is an integral homology equivalence as long as is path-connected.
The point of this theorem is that the right-hand side may be computed: the rational cohomology of may be described compactly as , the free graded-commutative algebra on the reduced cohomology of .
Proof sketch.
Following [12], and keeping track of the role that the maps to play, one may show that . Furthermore, under this equivalence the composition
is weakly homotopic to by the analogue (with maps to ) of the discussion in Section 5.3 of [12].
Now we claim that the natural map
is an integral homology equivalence as long as is path-connected. This may be proved following [12] and also using ideas from [13]. The crucial point is to consider the category where objects contain the origin and morphisms contain the interval on which the function is constant. The inclusion may be seen to induce an equivalence on classifying spaces as in [13, Lemma 4.6]. One then considers the subcategory having only object and in which the morphisms are required to be connected. We claim that this inclusion induces an equivalence on classifying spaces. To make morphisms connected use a move similar to that of [12, Lemma 4.24] which connects an arbitrary path component to the standard stick (this requires to be path-connected). To reduce to a single object use a move similar to that of [13, Section 4] to make objects consist of a single point, then isotope them into standard position as in [13, Proposition 4.26].
The category is therefore a monoid, and, as we can slide edges along the standard interval , it is a homotopy commutative monoid. We may thus apply the group-completion theorem to it, showing that
is a homology isomorphism, where the homotopy colimit is formed by left multiplication with a connected graph of genus 1. By precomposing with a morphism given by an interval, this map is easily compared with that in the statement of the theorem. ∎
The main technical result we will require is the following homological stability theorem for the spaces . We will deduce it from (two) arguments of Cohen–Madsen in the analogous situation of surfaces with maps to a background space.
Theorem 4.2.
Suppose that is simply-connected.
-
(i)
The map induces a homology isomorphism in degrees .
-
(ii)
The map induces an isomorphism on homology with -module coefficients in degrees .
Proof.
For part (i) we follow the argument of Cohen and Madsen [6]. There is a map of homotopy fibre sequences
so it is enough to show that is part of a polynomial coefficient system of degree . In this case the map of Serre spectral sequences will induce an isomorphism on for and an epimorphism for ; in particular it induces an isomorphism for and an epimorphism for . It follows from the spectral sequence comparison theorem that the map in an isomorphism in degrees and an epimorphism in degrees .
We define a coefficient system on objects by
and on a morphism we use precomposition by
This defines a coefficient system, and since is split surjective the stabilisation maps are all split injective. The coefficient system agrees with the constant coefficient system , as has been assumed to be simply-connected: in particular it is polynomial of degree 0.
Let us suppose for an induction that has degree for all . Consider the homotopy fibre sequence
where the map induces on th homology. As is split injective, the Serre spectral sequence for this fibration (which is over a path-connected base) collapses, and we find that has a filtration with associated graded
Each is polynomial of degree so is too; as degree is preserved under extensions it follows that has degree , hence has degree .
For part (ii), we follow a different argument of Cohen and Madsen [7]. We have a diagram
in which the row and column are fibrations, and the map is the adjoint to the map of spectra given by the composition of the Becker–Gottlieb transfer for the map followed by the map given by evaluating the map to at the marked point.
By part (i), Theorem 4.1, and the analogue with maps to of the discussion in Section 5.3 of [12], the composition is an isomorphism in degrees , and in particular is surjective on homology in this range. It follows from Leray–Hirsch that the map
is also an isomorphism in this range.
Now we have the commutative diagram
where denotes the homology suspension. The transfer map is (split) injective with -module coefficients, so in degrees the map is injective too. Hence is an isomorphism in this range, so is too. ∎
5. Proof of the main theorems
5.1. Recollections on Schur–Weyl duality
We shall need a small amount of representation theory, but nothing beyond e.g. the first few parts of Chapter 9 of [22]. Recall that for a vector space (over ) and a partition we have defined as . If is finite-dimensional then as the Specht modules give a complete set of irreducible representations of we may write
as -modules. The are non-zero as long as , in which case they are irreducible -modules.
5.2. Labeled partitions
Let us fix a sequence of -vector spaces , and consider
If we consider this as a functor landing in graded vector spaces, by letting have degree 1, then the homogenous pieces of are each polynomial functors of (see [22, §9.7] for this notion). Thus the -module is independent of as long as . We wish to identify this -module.
Choose bases for the vector spaces , and let be the set of partitions of equipped with a labeling of each part of size with an element . For each such datum there is a map of -modules
If then , giving a -equivariant map
Proposition 5.1.
The map is an isomorphism as long as .
Proof.
Let us call an orbit of acting on a type of partition. It consists of a partition along with an unordered list of labels in for the parts of size , and we write for this orbit.
The source of splits as a direct sum of (cyclic) modules one for each type of partition. For a type of partition and a , let us write and for the number of parts labeled by . The target of splits as a direct sum of terms
one for each type of partition. The map restricts to a -module map
and it is enough to show that each of these is an isomorphism.
As -representations we have a surjection
which is split by the standard symmetrisers. Furthermore if we choose a then we obtain an isomorphism
(where these terms must be suitably permuted to be put in the right form).
In total this identifies with a summand of , which by the first fundamental theorem of invariant theory for is a free left -module generated by the identity map of (as long as ). Thus the -module may be identified with the cyclic left submodule of generated by the idempotent , where is the stabiliser of .
Under this identification we have , so is surjective. On the other hand, we claim that the submodule is isomorphic to , and hence has dimension the size of the orbit . This implies that is an isomorphism. To prove the claim, note that the surjective module map sends to 0, so gives a surjection . On the other hand the module map
is well-defined and surjective. ∎
5.3. Proof of Theorem A
We have already proved Theorem A (i) in Corollary 3.6. For Theorem A (ii), first choose a functorial model for Eilenberg–MacLane spaces , then fix a finite-dimensional -vector space and consider the fibration
and its associated Serre spectral sequence with -coefficients
(5.1) |
The action of on , and hence on the fibration above, make this into a spectral sequence of -modules.
The proof of the following key lemma is close to an argument communicated to the author by Søren Galatius to prove [23, Conjecture C]. It was his argument that led us to think along the lines necessary to prove Theorem A.
Lemma 5.2.
The spectral sequence (5.1) collapses.
Proof.
The action of the scalars on makes it a spectral sequence of -modules, and the action of on the -vector space is with weight , i.e. acts by scalar multiplication by . As has characteristic zero, distinct weights make into distinct irreducible -modules. Thus there can be no -module maps between different rows of this spectral sequence, so it collapses. ∎
Furthermore, this argument identifies with the subspace of on which acts with weight .
Lemma 5.3.
If then is zero unless , in which case it is isomorphic to the degree part of .
Proof.
By Theorem 4.1 and Theorem 4.2, the map
is an isomorphism on cohomology in degrees . With coefficients of characteristic zero, , and taking the free infinite loop space has the effect of forming the free graded-commutative algebra, so in this case the symmetric algebra. The action of on is with weight precisely in degree . ∎
Corollary 5.4.
If then is zero unless , in which case it is isomorphic to the degree part of .
This proves the vanishing part of Theorem A (ii). We shall now use the fact that the identifications made so far are functorial in , and so are in particular -equivariant and can be decomposed into irreducible -modules. The following is a standard consequence of Schur–Weyl duality, but we explain its proof anyway.
Lemma 5.5.
As a -representation,
where denotes the conjugate (i.e. transpose) Young diagram to .
Proof.
The left-hand side is the -invariants in the -module . By our definition of Schur functors we have an isomorphism of -modules
where is the Specht module. Similarly , so tensoring them together and taking -invariants, using , the result follows. ∎
Putting the above together, we have an isomorphism of -modules
Choosing to be at least -dimensional, the are then distinct non-zero irreducible -modules, so by Schur’s lemma applying gives
as long as . Using , we obtain
and using that we can write the right-hand side as
Along with Proposition 5.1 this finishes the proof of Theorem A.
Remark 5.6.
We could have used instead of . In this case the argument goes through, the analogue of Lemma 5.5 is , and the result obtained is
whenever and , where denotes the free graded-commutative algebra.
A consequence of this is that the multiplicity of in is the same as the multiplicity of in , which does not seem obvious to the author.
5.4. Proof of Theorem B
We will make use of the following lemma, which follows from Kawazumi [18, Theorem 7.1]. It also follows from general principles: the Becker–Gottlieb transfer with local coefficients.
Lemma 5.7.
For any -module , the map
is split injective, and
Theorem B (i) and the first part of Theorem B (ii) follows immediately from this lemma, as it implies that and are summands of and respectively, so vanish under the stated assumptions.
It remains to prove the second part of Theorem B (ii). To do this, we again consider a finite-dimensional -vector space and consider the diagram
where the row is a (split) fibration and the column is a (trivial) fibration. This gives a spectral sequence
(5.2) |
Rather than a weight argument, which is no longer conclusive, we will deduce the degeneration of this spectral sequence from the vanishing results already established, namely that in degrees if . By Lemma 5.5 this implies that as long as and , so in bidegrees such that the -page of the spectral sequence (5.2) is
where the th summand lies in . In particular, all summands in this range lie in even total degree, and so there can be no differentials.
Lemma 5.8.
There is an isomorphism of graded -representations
and is independent of for .
Proof.
Writing using Lemma 5.5, and then identifying the abutment of the spectral sequence (5.2) as in Lemma 5.3, in total degree we obtain
as long as .
The right-hand side is independent of . For a partition we find that
is independent of as long as . By induction we may assume that all terms with are also independent of , leaving just the terms with , which gives by Schur’s lemma. This proves the second part.
Passing to the limit and summing over all terms above gives an isomorphism of graded -representations
Because the graded -representation
is the trivial representation in grading zero, the isomorphism in the lemma can be established by induction on degree similarly to the argument above. More conceptually, the graded virtual representation is invertible under , as it is the trivial representation in grading zero, and cancelling it from both sides gives the required isomorphism. ∎
5.5. Proof of Theorem C
6. Integral and torsion calculations
The general technique we have been using is not confined to rational coefficients. To give an example of how it may be used more generally, we shall now develop a “tame” strengthening of Theorem A, namely Theorem E (ii). Let us first explain how Theorem E (ii) implies Theorem E (i).
Proof of Theorem E (i).
By Theorem A we have that and all other rational cohomology groups vanish. By Theorem E (ii), for each prime number , is a free -module, so by the calculation above is a free module on a single generator in degree 1. Hence is , as it becomes a free -module of rank 1 when localised at any prime . Choosing a generator for this group gives a map
which must be an isomorphism, as it is so when localised at every prime. ∎
In order to prove Theorem E (ii) we first revisit some of the techniques we have used earlier and develop them in the -local rather than rational setting.
Let be fixed and be a prime number such that . Recall that to a partition and a tableau of shape there is an associated Young symmetriser , which satisfies . As is a -local unit we may form the element and this is an idempotent. If then , so there is a conjugacy class of idempotent associated to each . We write for the idempotent associated to the canonical tableau for . The Specht module is defined by , and by analogy we define the -local Specht module by . It is a free -module of rank , and is indecomposable as a -module (as ). These idempotents satisfy if and are tableaux of different shapes. Furthermore we have a decomposition into primitive idempotents
If is a -module then is a -module, and we define the -local Schur functor by . There is then a natural map
of -modules given by . For an inverse, define
where for a tableau of shape we write . This establishes the Schur–Weyl decomposition in this setting.
We now require a partial analogue of the notion of weights which we used in the proof of Lemma 5.2. For a -module and an integer , write for the -module which is the same as a -module and on which acts as scalar multiplication by ; say that it is a module which is pure of weight .
Lemma 6.1.
If and are -modules then
if .
Proof.
By the natural isomorphisms it is enough to establish the lemma for . In this case consider the functor
When this satisfies
By [1, Lemma 2.12] the of the numbers over all integers coprime to is itself coprime to as long as , and hence under this assumption on . The functor is exact, so taking derived functors of shows that the claim in the lemma holds for , and hence for any free -module. The claim in general follows by resolving by free modules. ∎
We now give the proof of Theorem E (ii), which states that under our assumptions on and , is a free -module.
Proof of Theorem E (ii).
Let be an odd integer, and let be a choice of model for the -local -sphere. For each unit we may find a map inducing multiplication by on , and these satisfy . Set , and consider the space , the Serre fibration
(6.1) |
and its associated Serre spectral sequence.
Recall that the -cohomology ring of is the divided power algebra on the class obtained by looping a generator of . This identifies the -cohomology ring of with , so, taking the limit , there is a spectral sequence
It follows from the results of [5, p. 40] that the map
from the free commutative algebra, induced by the map , is an isomorphism on homology in degrees . By the Künneth theorem we have
which is free as a -module. This shows that is a free -module in degrees , and shows that it is supported in degrees divisible by . Using Galatius’ theorem and the Künneth theorem again we may therefore identify the target of the spectral sequence with in a range of degrees. In total degrees the spectral sequence takes the form
and is a spectral sequence of -modules.
For the homotopy equivalence induces a map of spectral sequences by functoriality, making it into a spectral sequence of -modules. The induced map on is given by scalar multiplication by , so the st row of the spectral sequence is pure of weight . Furthermore, the fibration (6.1) has a section so there are no differentials entering the bottom row, and this row splits off the filtration. The rows have different weights which differ by at most , so by Lemma 6.1 there are no differentials in total degree and the associated filtration of splits as -modules.
The map induced by on is multiplication by , so this is pure of weight . It follows that the st row of the spectral sequence may be identified in a range of degrees with and so is a free -module. The -module contains as a summand in degree , so contains as a summand in this degree too. Thus the st row of the spectral sequence contains as a summand, so this is a projective -module in degrees , so in all degrees as and was arbitrary. Finally, as is a connected graded algebra over the local ring , projective graded modules which are finitely-generated in each degree are free. ∎
Appendix A An integral calculation for
It seems reasonable to suppose that Theorem 4.2 (ii) holds with integral and not just -module coefficients, that is, that
Conjecture A.
The map induces an isomorphism on homology in degrees .
Putting this together with Theorem 4.1, it follows that is an isomorphism on integral homology in degrees . Assuming this conjecture, we may make the following calculation.
Proposition A.1.
We have
We emphasise that this proposition holds without assuming Conjecture A: it follows from Theorem B that these groups are torsion, Satoh [26] has computed that for , and one easily computes that . Our purpose here is to give another proof of this proposition using Conjecture A.
This proposition should be contrasted with with a theorem of Bridson and Vogtmann [2, Theorem B], who show that the extension
is non-trivial for all , and hence gives a non-trivial class . Our calculation along with the result of Bridson and Vogtmann [2] that their class remains non-trivial in the group for any not coprime to implies that the class generates as long as .
Proof of Proposition A.1 assuming Conjecture A.
Consider the diagram
where the row and column are fibrations. Note that is a trivial fibration and is split via the inclusion of the constant maps, and there is an inclusion of the graphs with constant maps to . The Leray–Serre spectral sequence for the horizontal fibration is
(A.1) |
where is the canonical class in , so has bidegree .
We first claim that the map
has image on second cohomology. By our conjecture, it is enough to prove this after composing with the map as long as . Up to translation of components, this map is given by the Becker–Gottlieb transfer for the trivial graph bundle over composed with projection to . By standard properties of the transfer, this is times the standard inclusion, which on second cohomology induces multiplication by , as required.
This describes the edge homomorphism of the spectral sequence (A.1). We now use the homotopy equivalence
and that is 3-connected (by the Freudenthal suspension theorem) to describe the cohomology of in low degrees. Using that is for , it follows that is for . As the spectral sequence (A.1) converges to zero for positive Leray filtration in total degree , the differential must be onto (so is cyclic) and the kernel is , so . On the other hand, in total degree we see and , and it converges to , so observing the direction of the Leray filtration we see that . ∎
Appendix B Mapping class groups of surfaces
Let denote the mapping class group of a surface of genus , and consider as a -module. In this case Poincaré duality gives . Looijenga has already computed in the stable range [20], but we wish to explain here how the computation may also be performed using the techniques of this paper.
Cohen and Madsen [6] have introduced spaces of surfaces diffeomorphic to equipped with a map to , and have identified the stable (co)homology of these spaces with that of the infinite loop space as long as is simply-connected. Furthermore, there is a fibration sequence
so an associated Serre spectral sequence.
Considering , the argument of Section 5.4 goes through without change and identifies
with
in a range of degrees increasing with , where is the graded vector space which is in degrees and zero otherwise. This may be written as
and the first term is in the stable range, so we obtain
in a range of degrees increasing with .
It follows that each is a free -module in the stable range. To describe the space of module generators we can proceed as in the proof of Theorem A, and hence identify as a graded -module with
Proposition 5.1 identifies the space of -module homomorphisms with the permutation module on the set of the following data: a partition of , a labeling of each part of size 1 in the set , and a labeling of each part of size in the set . Such a datum is given grading plus the sum of the degrees of the labels (which are given by their subscripts).
Example B.1.
When we find that is a free -module with generators in degrees .
When the partition may be labeled by , and the partition may have each part labeled by . As a -set the action is trivial on the first type of elements, and the second type form trivial orbits
and free orbits
with . Thus has -module generators given by the multiplicities of the trivial representation in the indicated permutation module, so in degrees . Similarly has -module generators given by the multiplicities of the sign representation in the indicated permutation module, so in degrees .
By considering the analogous space of surfaces of genus with one boundary equipped with maps to which map the boundary to the basepoint, one identifies with
in a range of degrees increasing with , from which may be computed as above. There is a similar description in terms of partitions, the only difference being that now pieces of a partition of size 1 may also be labeled by . For example, one finds that is a free -module with generators in degrees . This description of (in terms of labelled partitions) has previously been obtained by Kawazumi [19], cf. Theorem 1.B of that paper, even with integral as opposed to just rational coefficients. However Kawazumi does not establish this description as -modules, because his method of proof breaks the -symmetry.
Appendix C Tables
The result of Theorems A and B for is compiled in the tables below, which was computed with the SchurRings package for Macaulay2 [14], and checked via Theorem C and the table at the end of Section 4.3 of [21].
1 | 1 | 2 | 0 | 3 | 1 | 0 | 5 | 2 | 2 | 0 | 0 |
7 | 5 | 4 | 0 | 1 | 0 | 0 |
11 | 8 | 10 | 2 | 1 | 2 | 2 | 0 | 0 | 0 | 0 |
1 | 0 | 1 | 0 | 1 | 0 | 0 | 2 | 0 | 1 | 0 | 0 |
2 | 1 | 1 | 0 | 0 | 0 | 0 |
4 | 1 | 3 | 0 | 0 | 0 | 1 | 0 | 0 | 0 | 0 |
References
- [1] J. F. Adams. On the groups . II. Topology, 3:137–171, 1965.
- [2] Martin R. Bridson and Karen Vogtmann. Abelian covers of graphs and maps between outer automorphism groups of free groups. Math. Ann., 353(4):1069–1102, 2012.
- [3] Kenneth S. Brown. Cohomology of groups, volume 87 of Graduate Texts in Mathematics. Springer-Verlag, New York, 1994. Corrected reprint of the 1982 original.
- [4] Henri Cartan and Samuel Eilenberg. Homological algebra. Princeton University Press, Princeton, N. J., 1956.
- [5] Frederick R. Cohen, Thomas J. Lada, and J. Peter May. The homology of iterated loop spaces. Springer-Verlag, Berlin, 1976. Lecture Notes in Mathematics, Vol. 533.
- [6] Ralph Cohen and Ib Madsen. Surfaces in a background space and the homology of mapping class group. Proc. Symp. Pure Math., 80(1):43–76, 2009.
- [7] Ralph Cohen and Ib Madsen. Stability for closed surfaces in a background space. Homology Homotopy Appl., 13(2):301–313, 2011.
- [8] Aurélien Djament. Décomposition de Hodge pour l’homologie stable des groupes d’automorphismes des groupes libres. arXiv:1510.03546, 2015.
- [9] Aurélien Djament and Christine Vespa. Sur l’homologie des groupes d’automorphismes des groupes libres à coefficients polynomiaux. Comment. Math. Helv., 90(1):33–58, 2015.
- [10] W. G. Dwyer. Twisted homological stability for general linear groups. Ann. of Math. (2), 111(2):239–251, 1980.
- [11] Johannes Ebert and Oscar Randal-Williams. Stable cohomology of the universal Picard varieties and the extended mapping class group. Doc. Math., 17:417–450, 2012.
- [12] Søren Galatius. Stable homology of automorphism groups of free groups. Ann. of Math. (2), 173(2):705–768, 2011.
- [13] Søren Galatius and Oscar Randal-Williams. Monoids of moduli spaces of manifolds. Geom. Topol., 14(3):1243–1302, 2010.
- [14] Daniel R. Grayson and Michael E. Stillman. Macaulay2, a software system for research in algebraic geometry. Available at http://www.math.uiuc.edu/Macaulay2.
- [15] Allen Hatcher and Karen Vogtmann. Cerf theory for graphs. J. London Math. Soc. (2), 58(3):633–655, 1998.
- [16] Allen Hatcher and Karen Vogtmann. Homology stability for outer automorphism groups of free groups. Algebr. Geom. Topol., 4:1253–1272 (electronic), 2004.
- [17] Nikolai V. Ivanov. On the homology stability for Teichmüller modular groups: closed surfaces and twisted coefficients. In Mapping class groups and moduli spaces of Riemann surfaces (Göttingen, 1991/Seattle, WA, 1991), volume 150 of Contemp. Math., pages 149–194. Amer. Math. Soc., Providence, RI, 1993.
- [18] Nariya Kawazumi. Cohomological aspects of Magnus expansions. arXiv:math/0505497, 2005.
- [19] Nariya Kawazumi. On the stable cohomology algebra of extended mapping class groups for surfaces. In Groups of diffeomorphisms, volume 52 of Adv. Stud. Pure Math., pages 383–400. Math. Soc. Japan, Tokyo, 2008.
- [20] Eduard Looijenga. Stable cohomology of the mapping class group with symplectic coefficients and of the universal Abel-Jacobi map. J. Algebraic Geom., 5(1):135–150, 1996.
- [21] L. Manivel. Gaussian maps and plethysm. In Algebraic geometry (Catania, 1993/Barcelona, 1994), volume 200 of Lecture Notes in Pure and Appl. Math., pages 91–117. Dekker, New York, 1998.
- [22] Claudio Procesi. Lie groups. Universitext. Springer, New York, 2007.
- [23] Oscar Randal-Williams. The stable cohomology of automorphisms of free groups with coefficients in the homology representation. arXiv:1012.1433, 2010.
- [24] Oscar Randal-Williams. The space of immersed surfaces in a manifold. Math. Proc. Cambridge Philos. Soc., 154(3):419–438, 2013.
- [25] Oscar Randal-Williams and Nathalie Wahl. Homological stability for automorphism groups. arXiv:1409.3541, 2014.
- [26] Takao Satoh. Twisted first homology groups of the automorphism group of a free group. J. Pure Appl. Algebra, 204(2):334–348, 2006.
- [27] Takao Satoh. Twisted second homology groups of the automorphism group of a free group. J. Pure Appl. Algebra, 211(2):547–565, 2007.
- [28] Christine Vespa. Extensions between functors from groups. arXiv:1511.03098, 2015.