Theorem 2.2 illustrates that the space of block diffeomorphisms \(\widetilde{\mathrm {Diff}}_{D^{2n}}(V_g)\) is closely related to the space \(\mathrm {\widetilde{hAut}}_{D^{2n}}(V_g,W_{g,1})\) of relative block homotopy automorphisms or, equivalently, to its non-block variant \(\mathrm {hAut}_{D^{2n}}(V_g,W_{g,1})\) (see Section 1.5). To access the homology of the classifying space of this space of homotopy automorphisms, one might try to study the Serre spectral sequence of the fibration sequence induced by taking components
$$\begin{aligned}\mathrm {BhAut}^{\mathrm {id}}_{D^{2n}}(V_g,W_{g,1})\longrightarrow \mathrm {BhAut}_{D^{2n}}(V_g,W_{g,1})\longrightarrow \mathrm {B}\pi _0\mathrm {hAut}_{D^{2n}}(V_g,W_{g,1})\end{aligned}$$
for which one ought to know at least the homology of the fibre as a module over the group \(\pi _0\mathrm {hAut}_{D^{2n}}(V_g,W_{g,1})\). This is what this section aims to compute—p-locally and in a range of degrees—by first calculating the p-local homotopy groups in a range using some tools from rational homotopy theory combined with an ad-hoc extension to the p-local setting tailored to our situation, and then pass from homotopy to homology groups.
Conventions on gradings
Essentially all objects in this section carry a \(\mathbf {Z}\)-grading, and we shall keep track of it throughout. For instance, we consider the (reduced) homology of a space X always with its natural grading, even if it is supported in a single degree.
We denote the k-fold suspension of a graded R-module A over some commutative ring R by \(s^kA\), the graded R-module whose degree k piece consists of R-module morphisms raising the degree by k by \(\mathrm {Hom}(A,B)\) for graded modules A and B, the graded R-dual of A by \(A^\vee :=\mathrm {Hom}(A,R[0])\) where R[0] is the base ring concentrated in degree zero, and the subspace of elements of strictly positive degrees by \(A^+\subset A\). For an ungraded R-module M we write M[k] for the graded R-module which is trivial in all degrees but k where it agrees with M. The graded tensor product \(A\otimes B\) is defined in the usual way. Note that \((s^kA)\otimes B=s^k(A\otimes B)=A\otimes (s^kB)\). The degreewise rationalisation or p-localisation of a graded \(\mathbf {Z}\)-module A is denoted by \(A_\mathbf {Q}\) or \(A_{(p)}\) respectively, and we view it as a graded \(\mathbf {Q}\)- respectively \(\mathbf {Z}_{(p)}\)-module.
Lie algebras and their derivations
We consider differential graded (short dg) Lie algebras over a commutative ring R. However, most of the dg Lie algebras which we shall encounter actually have trivial differential. Examples include the free graded Lie algebra \(\mathbf {L}(V)\) on a graded R-module A or the onefold shift of the homotopy groups \(\pi _{*+1}X\) of a based space X with its canonical Lie algebra structure over \(\mathbf {Z}\) given by the Whitehead bracket (except for a 2-torsion subtlety that will not play a role for us). Given a dg Lie algebra L, we write \([L,L]\subset L\) for the graded subspace generated by brackets. An important principle in this section is that the homotopy type of mapping spaces is closely related to certain chain complexes of f-derivations by which we mean the following: for a morphism \(f:(L,d_L)\rightarrow (L',d')\) of dg Lie algebras, an f-derivation of degree k is a linear map \(\theta :L\rightarrow L'\) that raises the degree by k and satisfies
$$\begin{aligned}\theta ([x,y])=[\theta (x),f(y)]+(-1)^{k|x|}[f(x),\theta (y].\end{aligned}$$
These derivations form the degree k piece of the chain complex \(\mathrm {Der}^f(L,L')\) of f-derivations over R whose differential is defined as \(d(\theta )=d_{L'}\theta -(-1)^{|\theta |}\theta d_L\), so it vanishes if both L and \(L'\) have trivial differential. Given a cycle \(\omega \in L\), we denote the subcomplex of f-derivations that vanish on \(\omega \) by \(\mathrm {Der}^f_{\omega }(L,L')\subset \mathrm {Der}^f(L,L')\). In the case \(L=L'\) and \(f=\mathrm {id}\), we abbreviate the complex of \(\mathrm {id}\)-derivations \(\mathrm {Der}^{\mathrm {id}}(L,L)\) by \(\mathrm {Der}(L)\).
Rational homotopy theory, Quillen style
Recall from [33] Quillen’s functor \(\lambda \), which assigns a simply connected based space X a dg Lie algebra \(\lambda (X)\) over the rationals, one of whose many properties is that it captures the rationalised homotopy Lie algebra of X via a natural isomorphism \(\mathrm {H}_*(\lambda (X))\cong \pi _{*+1}(X)_\mathbf {Q}\) of graded Lie algebras, where \(\mathrm {H}_*(\lambda (X))=\ker (d_{\lambda (X)})/\mathrm {im}(d_{\lambda (X)})\) is the homology Lie algebra of \(\lambda (X)\). A Lie model of X is a rational dg Lie algebra \(L_\mathbf {Q}^X\) quasi-isomorphic to \(\lambda (X)\). Such a model is called free if the underlying graded Lie algebra of \(L_\mathbf {Q}^X\) is isomorphic to a free graded Lie algebra \(\mathbf {L}(V)\) on a graded \(\mathbf {Q}\)-vector space V and minimal if it is free and has decomposable differential, i.e. \(d(L_\mathbf {Q}^X)\subset [L_\mathbf {Q}^X,L_\mathbf {Q}^X]\). Any simply connected based space has a minimal Lie model \(L_\mathbf {Q}^X\), unique up to (non-canonical) isomorphism, and a based map between such spaces \(f:X\rightarrow Y\) gives rise to a map \(f:L_\mathbf {Q}^X \rightarrow L_\mathbf {Q}^Y\) between their minimal models.
Derivations and mapping spaces
As mentioned earlier, the homotopy theory of mapping spaces is tightly connected to derivations of dg Lie algebras. In the rational setting, this is made precise for instance by a result of Lupton–Smith [30, Thm 3.1]. The version of their result we shall need is marginally stronger than stated in [30], but follows from the given proof in a straight-forward way (see also [8, Thm 3.6]).
Theorem 4.1
(Lupton–Smith) Let \(f:X\rightarrow Y\) be a map between simply connected finite based CW-complexes, with minimal Lie model \(f:L_\mathbf {Q}^X\rightarrow L_\mathbf {Q}^Y\). There is an isomorphism
$$\begin{aligned}\pi _*(\mathrm {Maps}_*(X,Y);f)_\mathbf {Q}\overset{\cong }{\longrightarrow }\mathrm {H}_*(\mathrm {Der}^{f}(L_\mathbf {Q}^X,L_\mathbf {Q}^Y))\end{aligned}$$
for \(*\ge 2\), which is natural in both X and Y. For X a co-H-space, this also holds for \(*=1\).
A p-local generalisation
From the point of view of Quillen’s approach to rational homotopy theory, the spaces we shall be applying Theorem 4.1 to are of the simplest nature possible: they are homotopy equivalent to boquets of equidimensional spheres. The minimal model of such a space \(X\simeq \vee ^gS^n\) for \(n\ge 2\) agrees with the free graded Lie algebra
$$\begin{aligned} L_\mathbf {Q}^{X}:=\mathbf {L}(s^{-1}H_\mathbf {Q}^X)\cong \pi _{*+1}X_\mathbf {Q}\quad \text {on}\quad H_\mathbf {Q}^X:=\widetilde{\mathrm {H}}_*(X;\mathbf {Q}),\end{aligned}$$
(36)
equipped with the trivial differential. Given a map between spaces of this kind, the induced map on minimals models is simply given by the induced map on rational homotopy groups. It is a consequence of the Hilton–Milnor theorem that, p-locally in small degrees with respect to p, the homotopy Lie algebra \(\pi _{*+1}X\) is free even before rationalisation. To make this precise, we abbreviate the integral and p-local analogue of (36) by
$$\begin{aligned}&L^X:=\mathbf {L}(s^{-1}H^{X})\text { and }L_{(p)}^X:=\mathbf {L}(s^{-1}H_{(p)}^{X})\text { where }\\&H^X:=\widetilde{\mathrm {H}}_*(X;\mathbf {Z})\text { and }H^X_{(p)}:=\widetilde{\mathrm {H}}_*(X;\mathbf {Z}_{(p)}).\end{aligned}$$
The inverse of the Hurewicz map \(\mathrm {H}_n(X;\mathbf {Z}_{(p)})\cong \pi _nX_{(p)}\) induces a map \(L_{(p)}^X\rightarrow \pi _{*+1}X_{(p)}\) of graded Lie algebras over \(\mathbf {Z}_{(p)}\), which turns out to be an isomorphism in a range of degrees.
Lemma 4.2
For an odd prime p and a based space X that is homotopy equivalent to \( \vee ^gS^n\) with \(n\ge 2\), the morphism
$$\begin{aligned}L_{(p)}^X\longrightarrow \pi _{*+1}X_{(p)}\end{aligned}$$
is an isomorphism on torsion free quotients. Moreover, the right hand side is torsion free in degrees \(*<2p-4+n\), so the map is an isomorphism in this range.
Proof
As a preparation to the proof, note that by specialising the Hilton–Milnor theorem to \(\vee ^gS^n\), we have an isomorphism
$$\begin{aligned} \textstyle {\pi _{i+1}(\vee ^gS^n)\cong \bigoplus _{\omega \in L_g}\pi _{i+1}(S^{l(\omega )(n-1)+1})}\end{aligned}$$
(37)
where \(L_g\) denotes a Hall basis for the free ungraded Lie algebra in g ordered generators and \(l(\omega )\) is the word-length of \(\omega \). Here the map \(\pi _{i+1}(S^{l(\omega )(n-1)+1})\rightarrow \pi _{i+1}(\vee ^gS^n)\) corresponding to \(\omega \in L_g\) is given by mapping a class \(x\in \pi _{i+1}(S^{l(\omega )(n-1)+1})\) to the composition \((\iota _\omega \circ x)\), where \(\iota _\omega \in \pi _{l(\omega )(n-1)+1}(\vee ^gS^{n})\) is the class obtained by taking Whitehead products of the canonical classes \(\iota _i \in \pi _{n}(\vee ^gS^n)\) for \(1\le i\le g\) represented by the inclusions of the summands as guided by the Lie word \(\omega \in L_g\). A proof can be extracted from [47]: combine XI.6.6 and the subsequent discussion with VII.2.6 and X.7.10.
To prove the asserted claim, we use that source and domain of the morphism in the statement are both degreewise finitely generated and that the rationalisation of this morphism agrees with (36), so to prove the first part of the claim, it suffices to show that all classes in \(\pi _{*+1}X_{(p)}\) of infinite order are in the image. From (37), we see every class in \(\pi _{k}X_{(p)}\) is a composition \((y \circ x)\) of some \(x\in \pi _kS^m_{(p)}\) with \(m\ge n\) and a class \(y\in \pi _mX_{(p)}\) in the image of the map in question. The group \(\pi _kS^m_{(p)}\) is finite unless \(k=m\), where it is generated by the identity, or \(m=2l\) and \(k=4l-1\), where it is is generated by \([\mathrm {id}_{S^{2l}},\mathrm {id}_{S^{2l}}]\), since this element has Hopf invariant 2 and we assumed p to be odd. As \(y\circ [\mathrm {id}_{S^{2l}},\mathrm {id}_{S^{2l}}]=[y,y]\) in \(\pi _{4l-1}X_{(p)}\) and the image of \(L_{(p)}^X\rightarrow \pi _{*+1}X_{(p)}\) is closed under taking brackets, this implies the first part of the claim. The second part follows from Serre’s result [38, p. 498, Prop. 5] that \(\pi _kS^m\) is p-torsion free for \(k-m<2p-3\) together with another application of (37). \(\square \)
As a result of Lemma 4.2, every map \(f:X\rightarrow Y\) between bouquets of equidimensional spheres induces a morphism \(f_*:L^X_{(p)}\rightarrow L^Y_{(p)}\) by taking torsion free quotients of p-local homotopy groups, so the following extension of Theorem 4.1 might not come as a surprise.
Proposition 4.3
For an odd prime p and a map \(f:X\rightarrow Y\) between based spaces \(X\simeq \vee ^g S^n\) and \(Y\simeq \vee ^h S^m\) with \(n,m\ge 2\) , the map of Theorem 4.1 fits into a commutative square
which is natural in X and Y and whose upper arrow is an isomorphism for \(*<2p-3-(n-m)\).
Remark 4.4
Dwyer’s tame homotopy theory [13] provides a p-local generalisation of Quillen’s rational homotopy theory for primes p that are just large enough with respect to the degree to prevent stable k-invariants from appearing. It is not unlikely that Theorem 4.1 could be generalised to this setting, but our layman extension Proposition 4.3 for bouquets of spheres suffices for the applications we have in mind.
Proof of Proposition 4.3
We begin with a twofold simplification of the statement. Firstly, the claimed naturality is automatic, since the vertical maps are evidently natural, the bottom map is natural by Theorem 4.1, and the right vertical map is injective, so it suffices to construct a top arrow with the desired properties for \(X=\vee ^gS^n\). Secondly, there is a commutative diagram
induced by restricting derivations to generators, which shows that it is enough to produce a dashed arrow making the diagram
commute. To do so, we consider the composition
$$\begin{aligned} \pi _*(\mathrm {Maps}_*(X,Y);f)\xrightarrow [\cong ]{(-f)_*}\pi _*(\mathrm {Maps}_*(X,Y);*)\cong \mathrm {Hom}(s^{-1}H^X,\pi _{*+1}Y)^+\end{aligned}$$
(39)
whose first isomorphism is given by acting with the inverse of f, using the loop space structure on \(\mathrm {Maps}_*(X,Y)\), and whose second isomorphism is induced by mapping a class in \( \pi _k(\mathrm {Maps}_*(X,Y);*)\) represented by a pointed map \(g:S^k\rightarrow \mathrm {Maps}_*(X,Y)\) to the composition
$$\begin{aligned}\mathrm {H}_n(X)\cong \mathrm {H}_{n+k}(S^k\wedge X)\cong \pi _{n+k}(S^k\wedge X)\overset{g_*}{\longrightarrow }\pi _{n+k}(Y),\end{aligned}$$
involving the suspension isomorphism, the inverse of the Hurewicz map, and the adjoint of g. Postcomposing (39) with the map given by p-localising and taking torsion free quotients results by Lemma 4.2 in a dashed map with the claimed connectivity property, so we are left to show that this choice does make the diagram (38) commute, i.e. that the rationalisation of (39) agrees with the bottom map of (38). The adjoint of a map \(h:S^k\rightarrow \mathrm {Maps}_*(X,Y)\) representing a class in \(\pi _*(\mathrm {Maps}_*(X,Y);f)\) forms the top arrow of the diagram
whose middle diagonal arrow is induced by h via the canonical homeomorphism \((S^k\times X)/(S^k\vee *)\cong S^k_+\wedge X\) and whose vertical equivalence is given as the composition
$$\begin{aligned}S^k_+\wedge X\xrightarrow {\mathrm {id}_{S^k_+}\wedge \nabla } S^k_+\wedge X\vee S^k_+ \wedge X\xrightarrow {c}X\vee S^k \wedge X\end{aligned}$$
using the co-H-space structure \(\nabla \) of X and the evident collapse map c. The map \((-f)_*(h)\) is the adjoint of a representative of the image of h under the first map in (39), so (40) commutes up to changing h within its class in \(\pi _k(\mathrm {Maps}_*(X,Y);f)\). We thus obtain a rational model for the top arrow in (40) as the composition
$$\begin{aligned} \begin{aligned}&\left( \mathbf {L}\left( s^{-1}H^X_{\mathbf {Q}}\oplus s^{-1+k}H_\mathbf {Q}^X\oplus s^{-1}H_\mathbf {Q}^{S^k}\right) ,d \right) \\ {}&\quad \longrightarrow L_\mathbf {Q}^{S^k\vee S^k\wedge X}\xrightarrow {\pi _{*+1}(f\vee (-f)_*(g))\otimes \mathbf {Q}} L_{\mathbf {Q}}^Y,\end{aligned}\end{aligned}$$
(41)
where the source is the Lie model of \(S^k\times X\) described in [30, Cor. 2.2], i.e. its differential d is trivial on \(s^{-1}H^X_{\mathbf {Q}}\oplus s^{-1}H_\mathbf {Q}^{S^k}\) and is on \(s^{-1+k}H_\mathbf {Q}^X\) given as
$$\begin{aligned} s^{-1+k}H_\mathbf {Q}^X\cong s^{-1}H^X_{\mathbf {Q}}\xrightarrow {(-1)^{k-1}[z,-]}\mathbf {L}\left( s^{-1}H^X_{\mathbf {Q}}\oplus s^{-1+k}H_\mathbf {Q}^X\oplus s^{-1}H_\mathbf {Q}^{S^k}\right) \end{aligned}$$
where the first isomorphism is the canonical identification as ungraded vector spaces induced by the identity and \(z\in s^{-1}H_\mathbf {Q}^{S^k}\) denotes the standard generator. The first map in the composition (41) takes the quotient by the dg Lie ideal generated by the subspace \(s^{-1}H_\mathbf {Q}^{S^k}\) and the second map is defined as indicated. Using this particular choice of rational model in the definition of the isomorphism of Theorem 4.1 in [30, p. 176–177], the image of the class defined by h under the bottom horizontal composition in (38) is precisely its image under (39) after rationalisation, so the claim follows. \(\square \)
p-local homotopy groups of \(\mathrm {BhAut}^{\mathrm {id}}_{D^{2n}}(V_g,W_{g,1})\)
The theory set up in the previous paragraphs will allow us to compute the p-local homotopy groups of the classifying space \(\mathrm {BhAut}^{\mathrm {id}}_{D^{2n}}(V_g,W_{g,1})\) of the identity component of the topological monoid of relative homotopy automorphisms as defined in Sect. 3.1 as a module over the group
$$\begin{aligned}\pi _1\mathrm {BhAut}_{D^{2n}}(V_g,W_{g,1})\cong \pi _0\mathrm {hAut}_{D^{2n}}(V_g,W_{g,1}).\end{aligned}$$
More generally, we will compute the p-local homotopy groups of the spaces participating in the fibration sequence (see Section 3.1 for the notation)
$$\begin{aligned} \mathrm {BhAut}_{\partial }(V_g)\longrightarrow \mathrm {BhAut}_{D^{2n}}(V_g,W_{g,1})\longrightarrow \mathrm {BhAut}^{\mathrm {ext}}_{\partial }(W_{g,1}) \end{aligned}$$
(42)
induced by restriction, together with the induced action of \(\pi _0\mathrm {hAut}_{D^{2n}} (V_g,W_{g,1})\). To state the answer (and give the proof), we adopt the notation of in the previous subsection for the three manifolds \(W_{g,1}\), \(V_g\), and \(\partial W_{g,1}\) involved, which are homotopy equivalent to bouquets of spheres. We do however omit the g-superscripts to increase readability, so write
$$\begin{aligned} H_{(p)}^W= & {} \widetilde{\mathrm {H}}_*(W_{g,1};\mathbf {Z}_{(p)}), L_{(p)}^W=\mathbf {L}(s^{-1}H_{(p)}^W),\\ H_{(p)}^V= & {} \widetilde{\mathrm {H}}_*(V_{g};\mathbf {Z}_{(p)}),\quad \text {and} \quad L_{(p)}^V=\mathbf {L}(s^{-1}H_{(p)}^V) \end{aligned}$$
and omit the (p)-subscripts to denote the integral analogues. Moreover, we generically write \(\iota \) for any combination of the inclusions
$$\begin{aligned}\partial W_{g,1}\subset W_{g,1}\subset W_g\subset V_g.\end{aligned}$$
Finally, we let \(\omega \in L_{W_{g,1}}\) be the class that represents the inclusion \(S^{2n-1}=\partial W_{g,1}\subset W_{g,1}\) of the boundary of \(W_{g,1}=\sharp ^g(S^n\times S^n)\backslash \mathrm {int}(D^{2n})\), i.e. the attaching map
$$\begin{aligned} \textstyle {\omega =\sum _{i=1}^{g}[e_i,f_i]\in \pi _{2n-1}W_{g,1}},\end{aligned}$$
(43)
where \(e_i,f_i\in \pi _nW_{g,1}\) correspond to the first respectively second \(S^n\)-summand in the ith summand of
$$\begin{aligned}W_{g,1}\cong \sharp ^{g}S^n\times S^n\backslash \mathrm {int}(D^{2n})\simeq \vee ^g (S^n\vee S^n).\end{aligned}$$
Note that the inclusion \(\partial W_{g,1}\subset V_g\) is trivial, since it factors over \(W_g=W_{g,1}\cup _{\partial W_{g,1}}D^{2n}\).
Remark 4.5
Note that \(H^{W}\) and \(H^{V}\) stand for the graded \(\mathbf {Z}\)-module given by the reduced homology of \(W_{g,1}\) and \(V_g\), which shall not be confused with the ungraded middle dimensional integral homology groups of these spaces that featured in Section 3.2.1 as \(H_{W_{g,1}}\) and \(H_{V_g}\).
Theorem 4.6
Let \(n\ge 2\) and p an odd prime.
-
(i)
The inclusion \(\pi _{0}\mathrm {Maps}_{\partial }(V_{g},V_g)\subset \pi _{0}\mathrm {hAut}_{\partial }(V_{g})\) is an equality. This group is abelian.
-
(ii)
In degrees \(0<*<2p-3-n\), the boundary map of the fibration (42) fits into a commutative diagram of graded \(\mathbf {Z}_{(p)}\)-modules with exact rows
which is \(\pi _0\mathrm {hAut}_{D^{2n}}(V_g,W_{g,1})\)-equivariant with respect to the action on the leftmost column induced by (42) and by the action through \(H^{W}\) and \(H^{V}\) on the other columns.
-
(iii)
Rationally, the conclusion of (ii) holds in all positive degrees.
A splitting of the canonical projection \(\iota _*:H_{W_{g,1}}\rightarrow H_{V_g}\) induces compatible splittings of the rightmost two columns of the diagram in Theorem 4.6 (ii), so the boundary map \(\partial \) of the fibration (42) is p-locally split surjective in a range and we conclude the following.
Corollary 4.7
Let \(n\ge 2\) and p an odd prime. In degrees \(0<*<2p-4-n\), the graded \(\mathbf {Z}_{(p)}[\pi _0\mathrm {hAut}_{D^{2n}}(V_g,W_{g,1})]\)-module \(\pi _{*+1}\mathrm {BhAut}_{D^{2n}}(V_g,W_{g,1})_{(p)}\) is isomorphic to the common kernel of the maps
$$\begin{aligned}&s^{-(2n-1)}H^{W}_{(p)}\otimes L_{(p)}^{W}\overset{\left[ -,-\right] }{\longrightarrow } s^{-(2n-2)}[L_{(p)}^{W}, L_{(p)}^{W}] \quad \text {and}\\&s^{-(2n-1)}H^{W}_{(p)}\otimes L_{(p)}^{W}\overset{\iota _*\otimes \iota _*}{\longrightarrow }s^{-(2n-1)}H^{V}_{(p)}\otimes L_{(p)}^{V}.\end{aligned}$$
Rationally, this holds in all positive degrees.
In particular, Theorem 4.6 and Corollary 4.7 imply that the \(\pi _0\mathrm {hAut}_{D^{2n}}(V_g,W_{g,1})\)-action on the p-local higher homotopy groups of the spaces participating in (42) factors in a range of degrees through the morphism (recall \(K_g=\ker (\mathrm {H}_n(W_{g,1})\rightarrow \mathrm {H}_n(V_g))\) from Section 3.2.1)
$$\begin{aligned}\pi _0\mathrm {hAut}_{D^{2n}}(V_g,W_{g,1})\longrightarrow \{\phi \in \mathrm {GL}(H_{W_{g,1}})\mid \phi (K_g)\subset K_g\}\end{aligned}$$
induced by the action on the homology of \(W_{g,1}\). During the proof of Theorem 4.6 and the preceding Lemma 4.8, we frequently use Proposition 4.3 to implicitly identify p-local homotopy groups of path components of pointed mapping spaces between bouquets of equidimensional spheres with derivations of free graded Lie algebras in a range of degrees. We denote by \(\mathrm {Maps}_*^f(X,Y)\) for a map of based spaces \(f:X\rightarrow Y\) the corresponding component of the mapping space, pointed by f. Reminding the reader of our notation for spaces of derivations in Section 4.2, we begin with the following lemma, whose first part is rationally due to Berglund and Madsen [8, Prop. 5.6].
Lemma 4.8
Let \(n\ge 2\) and p an odd prime.
-
(i)
In degrees \(*<2p-3-n\), the morphism induced by relaxing the boundary condition
$$\begin{aligned}\pi _*\mathrm {Maps}_\partial ^{\mathrm {id}}(W_{g,1},W_{g,1})_{(p)}\longrightarrow \pi _*\mathrm {Maps}_*^{\mathrm {id}}(W_{g,1},W_{g,1})_{(p)}\cong \mathrm {Der}(L_{(p)}^{W})^+\end{aligned}$$
is injective and has image \(\mathrm {Der}_{\omega }(L_{(p)}^{W})^+\subset \mathrm {Der}(L_{(p)}^{W})^+\).
-
(ii)
In the range \(*<2p-3-n\), the morphism induced by \(W_{g,1}\subset W_g\)
$$\begin{aligned}\pi _*\mathrm {Maps}_*^{\iota }(W_{g},V_g)_{(p)}\longrightarrow \pi _*\mathrm {Maps}_*^{\iota }(W_{g,1},V_g)_{(p)}\cong \mathrm {Der}^{\iota }(L_{(p)}^{W},L_{(p)}^{V})^+\end{aligned}$$
is injective and has image \(\mathrm {Der}^{\iota }_{\omega }(L_{(p)}^{W},L_{(p)}^{V})^+\subset \mathrm {Der}^{\iota }(L_{(p)}^{W},L_{(p)}^{V})^+\).
Proof
Restriction along the inclusion \(\partial W_{g,1}\subset W_{g,1}\) yields a fibration
$$\begin{aligned} \mathrm {Maps}_*(W_{g,1},W_{g,1})\longrightarrow \mathrm {Maps}_*(\partial W_{g,1},W_{g,1})\end{aligned}$$
(44)
whose fibre at \(\iota \) is \(\mathrm {Maps}_\partial (W_{g,1},W_{g,1})\). The induced maps on homotopy groups fits in the range \(0<*<2p-2-n\) into a diagram of the form
whose top square is provided by Proposition 4.3, so commutes. The bottom square is given as follows: the bottom right vertical map is the evaluation at the fundamental class
$$\begin{aligned}\in s^{-1}H_{(p)}^{\partial W_{g,1}}\cong \mathbf {Q}[2n-2],\end{aligned}$$
which factors as a composition of isomorphisms
$$\begin{aligned} \mathrm {Der}^{\iota }(L_{(p)}^{\partial W_{g,1}},L_{(p)}^{W})^+&\overset{\cong }{\longrightarrow }&\mathrm {Hom}(s^{-1}H_{(p)}^{\partial W_{g,1}}, L_{(p)}^{W})^+\\&\overset{\cong }{\longrightarrow }&\big (s^{-(2n-2)}L_{(p)}^{W}\big )^+=\big (s^{-(2n-2)}{[L_{(p)}^{W},L_{(p)}^{W}]}\big )^+\end{aligned}$$
where the first isomorphism restricts to generators, the second isomorphism evaluates at the fundamental class, and the final equality holds for degree reasons. The latter is because elements of degree \(>0\) in \(s^{-(2n-2)}L_{(p)}^{W}\) correspond to elements of degree \(>2n-2\) in \(L_{(p)}^{W}\), so are sums of brackets since this Lie algebra is generated in degree \(n-1\) as \(H_W\) is supported in degree n. The bottom left vertical map is the restriction to positive degrees of the map
$$\begin{aligned} \begin{aligned}\mathrm {Der}(L_{(p)}^{W})\overset{\cong }{\longrightarrow }\mathrm {Hom}(s^{-1}H_{(p)}^{W},L_{(p)}^{W})\cong&L_{(p)}^{W}\otimes (s^{-1}H_{(p)}^{W})^{\vee }\\\cong&s^{-(2n-1)}L_{(p)}^{W}\otimes H_{(p)}^{W}, \end{aligned} \end{aligned}$$
(46)
where the first isomorphism is given by restricting to generators, the second is the canonical one, and the third is induced by the intersection form on \(s^{-1}H_{W_{g,1}}\) (see Example B.1). By construction, the composition
$$\begin{aligned} \mathrm {Der}(L_{(p)}^{W})^+\longrightarrow \big (s^{-(2n-2)}[L_{(p)}^{W},L_{(p)}^{W}]\big )^+\end{aligned}$$
(47)
coincides with the evaluation at the class \(\omega \) that represents the inclusion \(\partial W_{g,1}\subset W_{g,1}\), so it follows from Lemma 1 that this square commutes up to a sign, since \(\omega =\sum _{i=1}^{g}[e_i,f_i]\in L_{W_{g,1}}\) agrees up to a sign with the element (76) from the appendix as \(e_i^{\#}= f_i\) and \(f_i^{\#}= e_i\) holds in the notation of the appendix up to a fixed sign depending on n (which does not play a role in the argument). As the bottom horizontal map is surjective as a consequence of the graded Jacobi identity, the middle horizontal arrow is surjective as well, and hence so is the top one. A consultation of the long exact sequence in homotopy groups induced by the fibration (44) thus proves (i) since the kernel of the middle horizontal arrow of the diagram is \(\mathrm {Der}_{\omega }(L_{(p)}^{W})\) as (47) is given by the evaluation at \(\omega \). This finishes the proof of (i). The proof of (ii) is completely analogous, based on the fibration sequence obtained by applying \(\mathrm {Maps}_*(-,V_g)\) to the cofibration sequence \(\partial W_{g,1}\rightarrow W_{g,1}\rightarrow W_g\) instead of (44). The bottom square of the diagram corresponding to (45) is now given by
whose vertical arrows are given by the composition
$$\begin{aligned} \mathrm {Der}^{\iota }(L_{(p)}^{\partial W_{g,1}},L_{(p)}^{V})^+&\overset{\cong }{\longrightarrow }&\mathrm {Hom}(s^{-1}H_{(p)}^{\partial W_{g,1}},L_{(p)}^{V})^+\\&\overset{\cong }{\longrightarrow }&\big (s^{-(2n-2)}L_{(p)}^{V}\big )^+=\big (s^{-(2n-2)}{[L_{(p)}^{V},L_{(p)}^{V}]}\big )^+\end{aligned}$$
and the restriction to elements of positive degrees of the composition
$$\begin{aligned} \begin{aligned} \mathrm {Der}^\iota (L_{(p)}^{W},L_{(p)}^{V})\xrightarrow {\cong }\mathrm {Hom}(s^{-1}H_{(p)}^{W},L_{(p)}^{V})\cong&L_{(p)}^{V}\otimes (s^{-1}H_{(p)}^{W})^{\vee }\\\cong&s^{-(2n-1)}L_{(p)}^{V}\otimes H_{(p)}^{W},\end{aligned}\end{aligned}$$
(49)
both completely analogous to the two compositions explained below (45). \(\square \)
Proof of Theorem 4.6
We consider the map of horizontal fibration sequences
where the top right horizontal arrow is induced by restriction and the rightmost vertical arrow is given by extending a selfmap of \(W_{g,1}\) relative to the boundary over the complement of \(W_{g,1}\subset W_{g}\) by the identity, followed by postcomposition with the inclusion \(W_{g}\subset V_g\). The induced morphism \(\pi _*\mathrm {Maps}^{\mathrm {id}}_*(V_g,V_g)\rightarrow \pi _*\mathrm {Maps}^{\iota }_*(W_g,V_g)\) is injective because its composition with the morphism \(\pi _*\mathrm {Maps}^{\iota }_*(W_g,V_g)\rightarrow \pi _*\mathrm {Maps}^{\iota }_*(W_{g,1},V_g)\) induced by restriction along the inclusion \(W_{g,1}\subset W_g\) is a retract since the composition
$$\begin{aligned}\vee ^g(S^n\vee S^n)\simeq W_{g,1}\subset W_g\subset V_g\simeq \vee ^gS^n\end{aligned}$$
is a homotopy retraction. From the long exact sequence in homotopy groups of the bottom fibration, we see that the monoid \(\pi _0\mathrm {Maps}_{\partial }(V_g,V_g)\) receives a surjection from \(\pi _1\mathrm {Maps}_*^{\iota }(W_g,V_g)\), and this is a monoid homomorphism as it agrees with the map induced on \(\pi _0(-)\) by the homotopy fibre inclusion of the fibre sequence of \(A_\infty \)-spaces
$$\begin{aligned}\big (\Omega \mathrm {Maps}_*^{\iota }(W_g,V_g) \simeq \mathrm {hofib}_{\mathrm {id}}(\mathrm {inc})\big )\longrightarrow \mathrm {Maps}_{\partial }(V_g,V_g)\overset{\mathrm {inc}}{\longrightarrow }\mathrm {Maps}_{*}(V_g,V_g).\end{aligned}$$
The group \(\pi _1\mathrm {Maps}_*^{\iota }(W_g,V_g)\) is abelian since we have
$$\begin{aligned} \pi _1\mathrm {Maps}_*^{\iota }(W_g,V_g)\cong [S^1\wedge W_g,V_g]_*\cong & {} [S^1\wedge (\vee ^{2g} S^n\vee S^{2n}),V_g]_*\\\cong & {} \pi _{n+1}(V_g)^{\oplus 2g}\oplus \pi _{2n+1}(V_g),\end{aligned}$$
using \(S^1\wedge W_g\simeq S^1\wedge (\vee ^{2g} S^n\vee S^{2n})\) due to the fact that the attaching map (43) of the top-dimensional cell in the usual CW-decomposition of \(W_g\) is a sum of Whitehead-brackets and thus nullhomotopic after suspension. Being surjected upon by an abelian group, the monoid \(\pi _0\mathrm {Maps}_{\partial }(V_g,V_g)\) is itself an abelian group and hence agrees with \(\pi _0\mathrm {hAut}_{\partial }(V_g)\), as claimed in (i). To prove (ii), we combine the injectivity we just observed with Lemma 4.8 (ii) and the long exact sequence of the bottom row in (50) to obtain a short exact sequence
$$\begin{aligned} 0\longrightarrow \mathrm {Der}(L_{(p)}^{V})^+&\overset{(-)\circ \iota _*}{\longrightarrow }&\mathrm {Der}^{\iota }_{\omega }(L_{(p)}^{W},L_{(p)}^{V})^+\nonumber \\\longrightarrow & {} \pi _{*-1}\mathrm {Maps}_\partial ^{\mathrm {id}}(V_g,V_g)_{(p)}\longrightarrow 0\end{aligned}$$
(51)
in the range \(*<2p-3-n\). Combining this with Lemma 4.8 (i), we see that the boundary map in the long exact sequence in homotopy groups of the upper fibration of (50) fits in degrees \(*<2p-3-n\) into a commutative diagram
where \(\pi \) is the quotient map and the right vertical map is induced by (51). To finish the proof of (ii), we thus need to show that the bottom composition of (52) fits as the left vertical arrow in a diagram as in (ii). To see this, we first combine the bottom square of (45) with the compatible square (48) to obtain a commutative diagram with exact rows
Next, writing
$$\begin{aligned}K:=\ker (\iota _*:H^W\rightarrow H^V),\end{aligned}$$
we note that there is a chain of natural isomorphisms
$$\begin{aligned} \mathrm {Der}(L_{(p)}^{V}){\cong }\mathrm {Hom}(H_{(p)}^{V},L_{(p)}^{V}){\cong } L_{(p)}^{V}{\otimes } (s^{-1}H_{(p)}^{V})^\vee {\cong } s^{-(2n-1)}L_{(p)}^{V}{\otimes } K_{(p)}\qquad \end{aligned}$$
(54)
defined analogously to (and compatible with) (49), using that the isomorphism \((H^{W})^\vee \cong H^{W}\) induced by the intersection form (neglecting grading shifts) sends \((H^{V})^\vee \subset (H^{W})^\vee \) to \(K\subset H^{W}\). Except for the equivariance claim, (ii) now follows by combining the chain of isomorphisms (54) with the diagrams (52)–(53) and the chain of isomorphisms
$$\begin{aligned}&\big (s^{-(2n-1)}L_{(p)}^{V}\otimes H_{(p)}^{W}\big )/\big ( s^{-(2n-1)}L_{(p)}^{V}\otimes K_{(p)} \big )\\&\quad \cong s^{-(2n-1)}L_{(p)}^{V}\otimes \big (H_{(p)}^{W}/ K_{(p)} \big )\cong s^{-(2n-1)}L_{(p)}^{V}\otimes H_{(p)}^{V}. \end{aligned}$$
This uses that the inclusion \(\mathrm {hAut}^{\mathrm {ext}}_\partial (W_{g,1})\subset \mathrm {Maps}_\partial (W_{g,1},W_{g,1})\) is 0-coconnected and that we have \(\mathrm {hAut}_\partial (V_g)=\mathrm {Maps}_\partial (V_g,V_g)\) by (i). To see the equivariance, note that all vertical maps in the diagram of (ii) are equivariant by construction. Since they are also surjective (see the discussion after the statement), it suffices to show that the top row is equivariant. This is clear for the second map in the top row, so we are left with showing equivariance of the first map
$$\begin{aligned} \pi _{*+1}\mathrm {BhAut}_\partial (W_{g,1})_{(p)}\longrightarrow s^{2n-1}H_{(p)}^{W}\otimes L_{(p)}^{W}.\end{aligned}$$
(55)
With respect to the canonical isomorphisms in positive degrees
$$\begin{aligned}\pi _{*+1}\mathrm {BhAut}_\partial (W_{g,1})_{(p)}\cong \pi _{*}\mathrm {hAut}_\partial (W_{g,1})_{(p)}\cong \pi _*\mathrm {Maps}_\partial (W_{g,1},W_{g,1})_{(p)},\end{aligned}$$
the action on the domain of (55) is induced by conjugation. Going through the proof, we see that (55) arises as a composition of the form
$$\begin{aligned} \pi _{*}\mathrm {Maps}^{\mathrm {id}}_\partial (W_{g,1},W_{g,1})_{(p)}\longrightarrow & {} \pi _{*}\mathrm {Maps}^{\mathrm {id}}_*(W_{g,1},W_{g,1})_{(p)}\\\longrightarrow & {} \mathrm {Der}(L_{(p)}^{W})\cong s^{2n-1}H_{(p)}^{W}\otimes L_{(p)}^{W}.\end{aligned}$$
The first map relaxes the boundary condition, which is equivariant. The second map is given by the isomorphism in Proposition 4.3, and its equivariance follows from the naturality part of that proposition. The third map is provided by the chain of isomorphisms (46), which is equivariant by the naturality of the intersection form. This finishes the proof of (ii). To see (iii), note that all restrictions on the degree in the proof of (ii) originated from the assumption on the degree in Proposition 4.3. This proposition holds rationally without that assumption, so the proof of 4.6 also applies to 4.6. \(\square \)
In a range of degrees, the particular shape of the p-local homotopy groups of the space \(\mathrm {BhAut}^{\mathrm {id}}_{D^{2n}}(V_g,W_{g,1})\) ensured by Corollary 4.7 allows us to pass from homotopy to homology groups, which is what we are actually interested in. In the following statement, we consider the module \(\mathrm {H}_n(W_{g,1};\mathbf {Z}_{(p)})\) as ungraded.
Corollary 4.9
For \(n\ge 2\), there is an injection of graded \(\pi _0\mathrm {hAut}_{D^{2n}} (V_g,W_{g,1})\)-modules
in degrees \(*< \min (2n-1,2p-3-n)\) for primes p.
Proof
We may assume \(p> 3\), since otherwise the claim has no content. As a result of Corollary 4.7, there is an injective map of graded \(\pi _0\mathrm {hAut}_{D^{2n}}(V_g,W_{g,1})\)-modules
in degrees \(0<*< 2p-4-n\). Using that we have
$$\begin{aligned}&[s^{-1}H^{W}_{(p)},s^{-1}H^{W}_{(p)}]\subset (s^{-1}H^{W}_{(p)})^{\otimes 2}\end{aligned}$$
by antisymmetrisation and that \(H^{W}_{(p)}=\widetilde{\mathrm {H}}_*(W_{g,1};\mathbf {Z}_{(p)})\) is concentrated in degree n, we obtain
$$\begin{aligned} s^{-(2n-1)}H^{W}_{(p)}\otimes \mathbf {L}(s^{-1}H^{W}_{(p)})\subset & {} s^{-(2n-1)}H^{W}_{(p)}\otimes (s^{-1}H^{W}_{(p)})^{\otimes 2}\\= & {} \big (\mathrm {H}_n(W_{g,1};\mathbf {Z}_{(p)})^{\otimes 3}\big )[n-1],\end{aligned}$$
in degrees \(*<2n-2\), so the claim holds for homotopy instead of homology groups. This leaves us with showing that the p-local Hurewicz homomorphism
$$\begin{aligned} \pi _{*}\mathrm {BhAut}^{\mathrm {id}}_{D^{2n}}(V_g,W_{g,1})_{(p)}\longrightarrow \widetilde{\mathrm {H}}_*(\mathrm {BhAut}^{\mathrm {id}}_{D^{2n}}(V_g,W_{g,1})_{(p)};\mathbf {Z}_{(p)}) \end{aligned}$$
is an isomorphism in degree \(*< m:=\min (2n-1,2p-3-n)\). Since submodules of free \(\mathbf {Z}_{(p)}\)-modules are free, it follows from the first part of the proof that n-truncation induces a p-locally m-connected map of the form
$$\begin{aligned}\mathrm {BhAut}^{\mathrm {id}}_{D^{2n}}(V_g,W_{g,1})\longrightarrow K(A,n)\end{aligned}$$
where A is a free \(\mathbf {Z}_{(p)}\)-module, so it suffices to show that K(A, n) has trivial \(\mathbf {Z}_{(p)}\)-homology in the range \(n<*<m\). Since A is free, it is enough to show that \(\mathrm {H}_*(K(\mathbf {Z}_{(p)},n);\mathbf {Z}_{(p)})\cong \mathrm {H}_*(K(\mathbf {Z},n);\mathbf {Z}_{(p)})\) vanishes in this range, which is certainly true rationally, so we may instead prove that \(\mathrm {H}^*(K(\mathbf {Z},n);\mathbf {F}_{p})\) vanishes for \(n+1<*<m+1\). As the natural map
$$\begin{aligned}\mathbf {HF}_p^*(\mathbf {HZ})={\lim }_n\mathrm {H}^{*+n}(K(\mathbf {Z},n);\mathbf {F}_p)\longrightarrow \mathrm {H}^{*+n}(K(\mathbf {Z},n);\mathbf {F}_p)\end{aligned}$$
is an isomorphism in degrees \(*<n\), this follows from showing that the spectrum cohomology \(\mathbf {HF}_p^*(\mathbf {HZ})\) vanishes in degrees \(0<*<\min (n,2p-2-2n)\). But \(\mathbf {HF}_p^*(\mathbf {HZ})\) is a quotient of the mod p Steenrod algebra \(\mathbf {HF}_p^*(\mathbf {HF}_p)\) by an ideal containing the Bockstein, so \(\mathbf {HF}_p^*(\mathbf {HZ})\) vanishes in degrees \(0<*<2p-2\) and we conclude the assertion. \(\square \)