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

Mapping class groups of exotic tori and actions by SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z})

Mauricio Bustamante Departamento de Matemáticas, Pontificia Universidad Católica de Chile [email protected] Manuel Krannich Department of Mathematics, Karlsruhe Institute of Technology, 76131 Karlsruhe, Germany [email protected] Alexander Kupers Department of Computer and Mathematical Sciences, University of Toronto Scarborough, 1265 Military Trail, Toronto, ON M1C 1A4, Canada [email protected]  and  Bena Tshishiku Department of Mathematics, Brown University [email protected]
Abstract.

We determine for which exotic tori 𝒯\mathcal{T} of dimension d4d\neq 4 the homomorphism from the group of isotopy classes of orientation-preserving diffeomorphisms of 𝒯\mathcal{T} to SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) given by the action on the first homology group is split surjective. As part of the proof we compute the mapping class group of all exotic tori 𝒯\mathcal{T} that are obtained from the standard torus by a connected sum with an exotic sphere. Moreover, we show that any nontrivial SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z})-action on 𝒯\mathcal{T} agrees on homology with the standard action, up to an automorphism of SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}). When combined, these results in particular show that many exotic tori do not admit any nontrivial differentiable action by SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}).

A homotopy dd-torus 𝒯\mathcal{T} is a dd-dimensional smooth manifold that is homotopy equivalent to the standard torus Td=×dS1T^{d}=\times^{d}S^{1} and hence also homeomorphic to it, by a known instance of the Borel conjecture; see [HW69] for d>4d>4, [FQ90, 11.5] for d=4d=4, [Wal68, 6.5] and the Poincaré conjecture for d=3d=3. If 𝒯\mathcal{T} is not diffeomorphic to the standard torus TdT^{d}, it is called exotic. For instance, given an exotic sphere Σ\Sigma of dimension kdk\leq d, the connected sum (TkΣ)×Tdk(T^{k}\sharp\Sigma)\times T^{d-k} is an exotic dd-torus.

One of the prominent features of the standard torus Td𝐑d/𝐙dT^{d}\cong\mathbf{R}^{d}/\mathbf{Z}^{d} is that it admits a faithful action SLd(𝐙)Diff+(Td)\mathrm{SL}_{d}(\mathbf{Z})\rightarrow\mathrm{Diff}^{+}(T^{d}) by SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) through orientation-preserving diffeomorphisms, induced by the linear action of SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) on 𝐑d\mathbf{R}^{d}. For a general homotopy dd-torus 𝒯\mathcal{T} one might thus wonder:

  1. (A)

    Is there a faithful action SLd(𝐙)Diff+(𝒯)\mathrm{SL}_{d}(\mathbf{Z})\rightarrow\mathrm{Diff}^{+}(\mathcal{T})? If not, is there even any nontrivial action?

As the SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z})-action on the standard torus splits the homomorphism Diff+(Td)SLd(𝐙)\mathrm{Diff}^{+}(T^{d})\rightarrow\mathrm{SL}_{d}(\mathbf{Z}) induced by the action on the first homology group H1(Td)π1(Td)𝐙d\mathrm{H}_{1}(T^{d})\cong\pi_{1}(T^{d})\cong\mathbf{Z}^{d}, it seems natural to approach Question (A) by first considering the following weaker question which is an instance of a high-dimensional version of a Nielsen realisation problem posed by Thurston [Kir97, Prob. 2.6]:

  1. (S)

    Is the homomorphism Diff+(𝒯)SLd(𝐙)\mathrm{Diff}^{+}(\mathcal{T})\rightarrow\mathrm{SL}_{d}(\mathbf{Z}) given by the action on H1(𝒯)\mathrm{H}_{1}(\mathcal{T}) split surjective?

This homomorphism factors through the mapping class group π0Diff+(𝒯)\pi_{0}\,\mathrm{Diff}^{+}(\mathcal{T}) of isotopy classes of orientation-preserving diffeomorphisms, so one can weaken the question further to:

  1. (S0)

    Is the homomorphism π0Diff+(𝒯)SLd(𝐙)\pi_{0}\,\mathrm{Diff}^{+}(\mathcal{T})\rightarrow\mathrm{SL}_{d}(\mathbf{Z}) given by the action on H1(𝒯)\mathrm{H}_{1}(\mathcal{T}) split surjective?

This work establishes several results regarding these three questions. Note that a positive answer to (S) implies positive answers to (A) and (S0). As part of our results, we

  • \bullet

    answer Question (S0) in all dimensions d4d\neq 4,

  • \bullet

    show that Questions (S) and (A) are in fact equivalent, and

  • \bullet

    conclude that for many exotic tori the answer to all three questions is negative.

In what follows, we describe these results and various extensions of them in more detail.

Splitting the homology action up to isotopy

Our first main result answers (S0) for d4d\neq 4:

Theorem A.

For a homotopy torus 𝒯\mathcal{T} of dimension d4d\neq 4, the morphism

π0Diff+(𝒯)SLd(𝐙)\pi_{0}\,\mathrm{Diff}^{+}(\mathcal{T})\longrightarrow\mathrm{SL}_{d}(\mathbf{Z})

induced by the action on H1(𝒯)\mathrm{H}_{1}(\mathcal{T}) is split surjective if and only if 𝒯\mathcal{T} is diffeomorphic to TdΣT^{d}\sharp\Sigma for a homotopy sphere ΣΘd\Sigma\in\Theta_{d} such that ηΣΘd+1\eta\cdot\Sigma\in\Theta_{d+1} is divisible by 22 in the abelian group Θd+1\Theta_{d+1}.

Here Θd\Theta_{d} is Kervaire–Milnor’s finite abelian group of homotopy dd-spheres [KM63] and ηΣΘd+1\eta\cdot\Sigma\in\Theta_{d+1} for ΣΘd\Sigma\in\Theta_{d} is the value of ηΣ\eta\otimes\Sigma under the Milnor–Munkres–Novikov pairing π1𝐒ΘdΘd+1\pi_{1}\,\mathbf{S}\otimes\Theta_{d}\to\Theta_{d+1} where ηπ1𝐒𝐙/2\eta\in\pi_{1}\,\mathbf{S}\cong\mathbf{Z}/2 is the generator of the first stable homotopy group of spheres (see [Bre67] for more on this pairing). The question whether ηΣΘd+1\eta\cdot\Sigma\in\Theta_{d+1} for a given ΣΘd\Sigma\in\Theta_{d} is divisible by 22 can in most instances be reduced to a problem in stable homotopy theory which can in turn be solved in many cases. This approach is discussed in Section 1.3, but to already illustrate its practicability at this point, we display in Table 1 below the first groups of homotopy spheres Θd\Theta_{d} together with the subgroups ΘdsplitΘd\smash{\Theta^{\mathrm{split}}_{d}\leq\Theta_{d}} of split spheres, i.e. those ΣΘd\Sigma\in\Theta_{d} for which ηΣ\eta\cdot\Sigma is divisible by 22, which is by A equivalent to π0Diff+(TdΣ)SLd(𝐙)\pi_{0}\,\mathrm{Diff}^{+}(T^{d}\sharp\Sigma)\rightarrow\mathrm{SL}_{d}(\mathbf{Z}) being split. Note that among the dimensions dd for which Θd\Theta_{d} is nontrivial, there are dimensions in which all spheres are split such as d=7d=7, dimensions in which none are split such as d=8d=8, as well as dimensions in which some but not all are split such as d=9d=9. In Section 1.3 we also explain why both cases—the sphere Σ\Sigma being split or not—occur for exotic spheres Σ\Sigma in infinitely many dimensions.

dd 6\leq 6 and 4\neq 4 77 88 99 1010 1111 1212 1313
Θd\Theta_{d} 0 𝐙/28\mathbf{Z}/28 𝐙/2\mathbf{Z}/2 (𝐙/2)2𝐙/2\ \ \ \ (\mathbf{Z}/2)^{\oplus 2}\oplus\mathbf{Z}/2 𝐙/6\mathbf{Z}/6 𝐙/992\mathbf{Z}/992 0 𝐙/3\mathbf{Z}/3
Θdsplit\ \ \,\,\Theta^{\mathrm{split}}_{d} 0 𝐙/28\mathbf{Z}/28 0 (𝐙/2)20(\mathbf{Z}/2)^{\oplus 2}\oplus 0 𝐙/6\mathbf{Z}/6 𝐙/992\mathbf{Z}/992 0 𝐙/3\mathbf{Z}/3
1414 1515 1616 1717 1818 1919
𝐙/2\mathbf{Z}/2 𝐙/2𝐙/8128\mathbf{Z}/2\oplus\mathbf{Z}/8128 𝐙/2\mathbf{Z}/2 (𝐙/2)3𝐙/2\ \ \ \,\,(\mathbf{Z}/2)^{\oplus 3}\oplus\mathbf{Z}/2 𝐙/8𝐙/2\mathbf{Z}/8\oplus\mathbf{Z}/2     𝐙/2𝐙/523264\mathbf{Z}/2\oplus\mathbf{Z}/523264
0 𝐙/2𝐙/8128\mathbf{Z}/2\oplus\mathbf{Z}/8128 0 (𝐙/2)30(\mathbf{Z}/2)^{\oplus 3}\oplus 0 𝐙/8𝐙/2\mathbf{Z}/8\oplus\mathbf{Z}/2     𝐙/2𝐙/523264\mathbf{Z}/2\oplus\mathbf{Z}/523264
Table 1. The groups Θd\Theta_{d} of homotopy dd-spheres for d19d\leq 19 together with the subgroups ΘdsplitΘd\smash{\Theta^{\mathrm{split}}_{d}\leq\Theta_{d}} of those ΣΘd\Sigma\in\Theta_{d} for which ηΣ\eta\cdot\Sigma is divisible by 22.

Actions of SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) on homotopy tori

Our second main result shows that all nontrivial SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z})-actions on homotopy tori agree on homology with the standard action up to an automorphism.

Theorem B.

Fix d3d\geq 3, a homotopy dd-torus 𝒯\mathcal{T}, and an automorphism group

G{Diff+(𝒯),Homeo+(𝒯)}.G\in\{\mathrm{Diff}^{+}(\mathcal{T}),\mathrm{Homeo}^{+}(\mathcal{T})\}.

Any homomorphism SLd(𝐙)G\mathrm{SL}_{d}(\mathbf{Z})\rightarrow G is either trivial or has the property that its postcomposition

SLd(𝐙)GSLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z})\longrightarrow G\longrightarrow\mathrm{SL}_{d}(\mathbf{Z})

with the action on H1(𝒯)\mathrm{H}_{1}(\mathcal{T}) is an automorphism. Moreover, if also d4,5d\neq 4,5, then the same holds when replacing GG by the group π0G\pi_{0}\,G of isotopy classes.

In particular, given any nontrivial homomorphism φ:SLd(𝐙)G\varphi\colon\mathrm{SL}_{d}(\mathbf{Z})\rightarrow G, we obtain a splitting of the action α:GSLd(𝐙)\alpha\colon G\rightarrow\mathrm{SL}_{d}(\mathbf{Z}) on first homology, given by φ(αφ)1\varphi\circ(\alpha\circ\varphi)^{-1}. Applying this to G=Diff+(𝒯)G=\mathrm{Diff}^{+}(\mathcal{T}) shows that the above questions (S) and (A) are in fact equivalent. Applying it to π0G=π0Diff+(𝒯)\pi_{0}\,G=\pi_{0}\,\mathrm{Diff}^{+}(\mathcal{T}) also shows that (S0) is equivalent to the following isotopy-analogue of (A).

  1. (A0)

    Is there a faithful action SLd(𝐙)π0Diff+(𝒯)\mathrm{SL}_{d}(\mathbf{Z})\rightarrow\pi_{0}\,\mathrm{Diff}^{+}(\mathcal{T})? If not, is there even any nontrivial action?

Combining these implications with A results in the following corollary which answers all questions (A), (S), (S0), (A0) in the negative for a large class of homotopy tori and partially answers Question 1.4 and Problem 1.5 in work of Bustamante and Tshishiku [BT21].

Corollary C.

Let 𝒯\mathcal{T} be a homotopy torus of dimension d4d\neq 4. If

  1. (i)

    𝒯\mathcal{T} is not diffeomorphic to a connected sum TdΣT^{d}\sharp\Sigma with ΣΘd\Sigma\in\Theta_{d}, or

  2. (ii)

    𝒯\mathcal{T} is diffeomorphic to TdΣT^{d}\sharp\Sigma for some ΣΘd\Sigma\in\Theta_{d} such that ηΣΘd+1\eta\cdot\Sigma\in\Theta_{d+1} is not divisible by 22,

then every homomorphism from SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) to Diff+(𝒯)\mathrm{Diff}^{+}(\mathcal{T}) or to π0Diff+(𝒯)\pi_{0}\,\mathrm{Diff}^{+}(\mathcal{T}) is trivial.

Remark (The Zimmer programme).

One motivation for considering Question (A) stems from the Zimmer programme, part of which studies actions of SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) on manifolds. For instance, it follows from a version of Zimmer’s conjecture, now a theorem due to Brown–Fisher–Hurtado [BFH20], that SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) does not act faithfully on smooth manifolds of dimension d2\leq d-2. For actions of SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) on dd-manifolds, there is a conjectural classification by Fisher–Melnick [FM22, Conjecture 3.6] which would imply that if SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) acts faithfully on a homotopy dd-torus 𝒯\mathcal{T}, then 𝒯\mathcal{T} is the standard torus. C implies this for a large class of homotopy tori.

Remark (Regularity).

Our results are phrased in terms of the group Diff+(𝒯)\mathrm{Diff}^{+}(\mathcal{T}) of CC^{\infty}-diffeomorphisms, but they also hold for the groups Diff+,k(𝒯)\mathrm{Diff}^{+,k}(\mathcal{T}) of CkC^{k}-diffeomorphisms for finite k1k\geq 1. For Theorems A and D, this follows from the isomorphism π0Diff+(𝒯)π0Diff+,k(𝒯)\pi_{0}\,\mathrm{Diff}^{+}(\mathcal{T})\cong\pi_{0}\,\mathrm{Diff}^{+,k}(\mathcal{T}). For B it follows from the observation that the statement for the group Homeo+(𝒯)\mathrm{Homeo}^{+}(\mathcal{T}) also implies the statement for all its subgroups. The deduction of C from Theorems A and B works the same way. In particular, this shows that homotopy tori as in C do not admit any C1C^{1}-action by SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}).

We conclude this introduction by explaining two results featuring in the proofs of A and B that may be of independent interest.

Mapping class groups of exotic tori

As an ingredient for the proof of A, we determine the mapping class groups π0Diff+(𝒯)\pi_{0}\,\mathrm{Diff}^{+}(\mathcal{T}) of exotic tori of the form 𝒯=TdΣ\mathcal{T}=T^{d}\sharp\Sigma for ΣΘd\Sigma\in\Theta_{d} in all dimensions d7d\geq 7 in terms of the known mapping class group π0Diff+(Td)\pi_{0}\,\mathrm{Diff}^{+}(T^{d}) of the standard torus. Note that Θd\Theta_{d} is trivial when d6d\leq 6 and d4d\neq 4, so in these cases there is nothing to show. To state the result, we first recall the previously known description of π0Diff+(Td)\pi_{0}\,\mathrm{Diff}^{+}(T^{d}). As mentioned above, the action of SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) on TdT^{d} induces a splitting of the action map π0Diff+(Td)SLd(𝐙)\pi_{0}\,\mathrm{Diff}^{+}(T^{d})\rightarrow\mathrm{SL}_{d}(\mathbf{Z}), so there is a semidirect product decomposition

π0Diff+(Td)=SLd(𝐙)π0TorDiff(Td)withπ0TorDiff(Td):-ker(π0Diff+(Td)SLd(𝐙)).\pi_{0}\,\mathrm{Diff}^{+}(T^{d})=\mathrm{SL}_{d}(\mathbf{Z})\ltimes\pi_{0}\,\mathrm{Tor}^{\mathrm{Diff}}(T^{d})\quad\text{with}\quad\pi_{0}\,\mathrm{Tor}^{\mathrm{Diff}}(T^{d})\coloneq\ker\big{(}\pi_{0}\,\mathrm{Diff}^{+}(T^{d})\rightarrow\mathrm{SL}_{d}(\mathbf{Z})\big{)}.

For d6d\geq 6, the kernel π0TorDiff(Td)\pi_{0}\,\mathrm{Tor}^{\mathrm{Diff}}(T^{d}) is abelian and isomorphic to the sum of 𝐙[SLd(𝐙)]\mathbf{Z}[\mathrm{SL}_{d}(\mathbf{Z})]-modules

(1) Ω:-(0jd(Λj𝐙d)Θdj+1)((Λd2𝐙d)𝐙/2)((𝐙/2)[𝐙d]/(𝐙/2)[1])C2\textstyle{\Omega\coloneq\Big{(}\bigoplus_{0\leq j\leq d}(\Lambda^{j}\mathbf{Z}^{d})\otimes\Theta_{d-j+1}\Big{)}\oplus\Big{(}(\Lambda^{d-2}\mathbf{Z}^{d})\otimes\mathbf{Z}/2\Big{)}\oplus\Big{(}(\mathbf{Z}/2)[\mathbf{Z}^{d}]/(\mathbf{Z}/2)[1]\Big{)}_{C_{2}}}

where SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) acts through the standard action on 𝐙d\mathbf{Z}^{d}, and ()C2(-)_{C_{2}} denotes the coinvariants with respect to the involution induced by multiplication by 1-1 on 𝐙d\mathbf{Z}^{d} (see [Hat78, Theorem 4.1, Remark (3) on p. 9]111[Hat78, Theorem 4.1] asserts that the computation of π0TorDiff(Td)\pi_{0}\,\mathrm{Tor}^{\mathrm{Diff}}(T^{d}) also holds for d=5d=5. However, this relies on a claim attributed to Igusa (see the middle of p. 7 loc.cit.) for which—to our knowledge—no proof has been provided so far. and [HS76, Theorem 2.5]). In addition to this description of π0TorDiff(Td)\pi_{0}\,\mathrm{Tor}^{\mathrm{Diff}}(T^{d}), our identification of π0Diff+(TdΣ)\pi_{0}\,\mathrm{Diff}^{+}(T^{d}\sharp\Sigma) involves the aforementioned homotopy sphere ηΣΘd+1\eta\cdot\Sigma\in\Theta_{d+1} and the unique nontrivial central extension

0𝐙/2SL¯d(𝐙)SLd(𝐙)00\longrightarrow\mathbf{Z}/2\longrightarrow\overline{\mathrm{SL}}_{d}(\mathbf{Z})\longrightarrow\mathrm{SL}_{d}(\mathbf{Z})\longrightarrow 0

of SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) by 𝐙/2\mathbf{Z}/2; see Section 2.1. Our result identifies the group π0Diff+(Td#Σ)\pi_{0}\,\mathrm{Diff}^{+}(T^{d}\#\Sigma) as a semidirect product of SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) or SL¯d(𝐙)\overline{\mathrm{SL}}_{d}(\mathbf{Z}) acting on a quotient of Ω\Omega by a nontrivial subgroup depending on Σ\Sigma which is contained in the summand Θd+1(𝐙dΘd)\Theta_{d+1}\oplus(\mathbf{Z}^{d}\otimes\Theta_{d}) of (1) corresponding to the terms j=0,1j=0,1.

Theorem D.

For a homotopy sphere ΣΘd\Sigma\in\Theta_{d} of dimension d7d\geq 7, there is an isomorphism

π0Diff+(Td#Σ){SL¯d(𝐙)[Ω/(ηΣ(𝐙dΣ)]if ηΣΘd+1is not divisible by 2SLd(𝐙)[Ω/(𝐙dΣ)]if ηΣΘd+1 is divisible by 2\pi_{0}\,\mathrm{Diff}^{+}(T^{d}\#\Sigma)\cong\begin{cases}\overline{\mathrm{SL}}_{d}(\mathbf{Z})\ltimes\Big{[}\Omega/\big{(}\langle\eta\cdot\Sigma\rangle\oplus(\mathbf{Z}^{d}\otimes\langle\Sigma\rangle\big{)}\Big{]}&\text{if }\eta\cdot\Sigma\in\Theta_{d+1}\text{is not divisible by }2\\ \hfil\mathrm{SL}_{d}(\mathbf{Z})\ltimes\Big{[}\Omega/\big{(}\mathbf{Z}^{d}\otimes\langle\Sigma\rangle\big{)}\Big{]}&\text{if }\eta\cdot\Sigma\in\Theta_{d+1}\text{ is divisible by }2\end{cases}

which is compatible with the homomorphisms to SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}).

In particular, this result shows that the mapping class group π0Diff+(Td#Σ)\pi_{0}\,\mathrm{Diff}^{+}(T^{d}\#\Sigma) for ΣΘd\Sigma\in\Theta_{d} is given by a quotient of SL¯d(𝐙)Ω\overline{\mathrm{SL}}_{d}(\mathbf{Z})\ltimes\Omega by a finite abelian subgroup which is always of order at least 22 and has order precisely 22 if and only if Σ\Sigma is the standard sphere, so from this mapping class point of view the standard torus admits “the most symmetries”, as one would expect.

Endomorphisms of SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z})

As an ingredient for the proof of B, we prove the following classification results for endomorphisms of SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) for d3d\geq 3:

Theorem E.

Fix d3d\geq 3. Every nontrivial endomorphism of SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) is an automorphism. Moreover, all automorphisms of SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) agree, up to postcomposition with a conjugation by an element in GLd(𝐙)\mathrm{GL}_{d}(\mathbf{Z}), with either the identity or the inverse-transpose automorphism.

Remark.

Some comments on E.

  1. (i)

    The proof is “elementary” in that it does neither rely on Margulis’ superrigidity or normal subgroup theorem, nor on the congruence subgroup property. Using these results, there are likely other proofs. The argument we give was hinted at by Ian Agol in a comment to a question on MathOverflow [Mat17] and sketched by Uri Bader in the case d=3d=3 as a response to the question (however this sketch has a small gap; see Remarks 4.3 and 4.8).

  2. (ii)

    For d=2d=2, the statement of E fails: consider the composition

    SL2(𝐙)H1(SL2(𝐙))𝐙/12SL2(𝐙)\mathrm{SL}_{2}(\mathbf{Z})\longrightarrow\mathrm{H}_{1}(\mathrm{SL}_{2}(\mathbf{Z}))\cong\mathbf{Z}/12\longrightarrow\mathrm{SL}_{2}(\mathbf{Z})

    where the first arrow is abelianisation and the second sends a generator to idSL2(𝐙)-\mathrm{id}\in\mathrm{SL}_{2}(\mathbf{Z}).

  3. (iii)

    The second part of E holds more generally; see [O’M66, Theorem A].

Acknowledgements

We would like to thank Wilberd van der Kallen for helpful comments. AK acknowledges the support of the Natural Sciences and Engineering Research Council of Canada (NSERC) [funding reference number 512156 and 512250]. AK was supported by an Alfred J. Sloan Research Fellowship. BT is supported by NSF grant DMS-2104346. MB is supported by ANID Fondecyt Iniciación en Investigación grant 11220330.

1. Collar twists

As preparation to the proof of Theorems A and D, we collect various results on a certain map SO(d)BDiff(M\int(Dd))\mathrm{SO}(d)\rightarrow\mathrm{BDiff}_{\partial}(M\backslash{\mathrm{int}(D^{d})}) defined by twisting a collar of the complement of an embedded dd-disc in a closed smooth dd-manifold MM. After explaining the construction, we discuss how this map behaves under taking products and connected sums, followed by some results on the collar twisting map for specific choices of MM, first homotopy spheres and then homotopy tori.

1.1. The collar twist

Given a closed connected oriented dd-dimensional manifold MM, we write

M:-M\int(Dd)M^{\circ}\coloneq M\backslash\mathrm{int}(D^{d})

for the complement of a fixed embedded disc DdMD^{d}\subset M that is compatible with the orientation (which is unique up to isotopy), and we write Diff(M)\mathrm{Diff}_{\partial}(M^{\circ}) for the group of diffeomorphisms of MM^{\circ} that fix a neighbourhood of the boundary sphere M=Sd1\partial M^{\circ}=S^{d-1} pointwise, equipped with the smooth topology. The latter is homotopy equivalent to the larger group DiffTM(M)\mathrm{Diff}_{T_{*}M}(M) of diffeomorphisms of MM that fix the centre of the disc M\ast\in M as well as the tangent space at this point. The group DiffTM(M)\mathrm{Diff}_{T_{*}M}(M) is the fibre of the fibration d:Diff+(M)GLd+(𝐑)\smash{d\colon\mathrm{Diff}_{\ast}^{+}(M)\rightarrow\mathrm{GL}^{+}_{d}(\mathbf{R})} assigning to a diffeomorphism that fixes \ast its (orientation-preserving) derivative at that point, so after delooping and using the equivalence GLd+(𝐑)SO(d)\smash{\mathrm{GL}^{+}_{d}(\mathbf{R})\simeq\mathrm{SO}(d)}, there is a homotopy fibration sequence

(2) BDiff(M)extBDiff+(M)𝑑BSO(d).\mathrm{BDiff}_{\partial}(M^{\circ})\overset{\mathrm{ext}}{\longrightarrow}\mathrm{BDiff}^{+}_{\ast}(M)\overset{d}{\longrightarrow}\mathrm{BSO}(d).

where ext\mathrm{ext} is induced by extending a diffeomorphism of MM^{\circ} to MM by the identity. The connecting map SO(d)ΩBSO(d)BDiff(M)\mathrm{SO}(d)\simeq\Omega\mathrm{BSO}(d)\rightarrow\mathrm{BDiff}_{\partial}(M^{\circ}) has the following geometric description: there is a homomorphism ΩSO(d)Diff([0,1]×Sd1)\Omega\mathrm{SO}(d)\rightarrow\mathrm{Diff}_{\partial}([0,1]\times S^{d-1}) which sends a smooth loop γΩSO(d)\gamma\in\Omega\mathrm{SO}(d) that is constant near the endpoints to the self-diffeomorphism of [0,1]×Sd1[0,1]\times S^{d-1} given by mapping (t,x)(t,x) to (t,γ(t)x)(t,\gamma(t)\cdot x), and a homomorphism ext:Diff([0,1]×Sd1)Diff(M)\mathrm{ext}\colon\mathrm{Diff}_{\partial}([0,1]\times S^{d-1})\rightarrow\mathrm{Diff}_{\partial}(M^{\circ}) induced by a choice of collar of the boundary sphere in MM^{\circ}. Delooping their composition gives a map

ΥM:SO(d)BDiff(M)\Upsilon_{M}\colon\mathrm{SO}(d)\longrightarrow\mathrm{BDiff}_{\partial}(M^{\circ})

that agrees with the aforementioned connecting map; see e.g. [Kra21, p. 9]. Following Section 3 of loc.cit., we call ΥM\Upsilon_{M} the collar twisting map of MM. This map is relevant to the study of the mapping class groups of MM and MM^{\circ}, since the sequence (2) induces an exact sequence of groups

(3) (π1SO(d){𝐙 if d=2𝐙/2 if d3)(ΥM)π0Diff(M)extπ0Diff+(M)0,\left(\pi_{1}\,\mathrm{SO}(d)\cong\begin{cases}\mathbf{Z}&\text{ if }d=2\\ \mathbf{Z}/2&\text{ if }d\geq 3\\ \end{cases}\right)\overset{(\Upsilon_{M})_{\ast}}{\longrightarrow}\pi_{0}\,\mathrm{Diff}_{\partial}(M^{\circ})\overset{\mathrm{ext}}{\longrightarrow}\pi_{0}\,\mathrm{Diff}^{+}_{\ast}(M)\longrightarrow 0,

so the second morphism in this sequence is an isomorphism if and only if the image

tM(ΥM)(1)π0Diff(M)t_{M}\coloneqq(\Upsilon_{M})_{\ast}(1)\in\pi_{0}\,\mathrm{Diff}_{\partial}(M^{\circ})

of the standard generator of the leftmost group under the first map (ΥM)(\Upsilon_{M})_{\ast} is trivial. We call this element the collar twist of MM. Note that the collar twist lies in the centre of π0Diff(M)\pi_{0}\,\mathrm{Diff}_{\partial}(M^{\circ}), because the image of the connecting map π2(X)π1(F)\pi_{2}(X)\rightarrow\pi_{1}(F) in the long exact sequence of homotopy groups for any fibration FEXF\rightarrow E\rightarrow X has this property. Alternatively, one could use that the collar twist is supported in a collar and that every diffeomorphism fixing boundary can be isotoped to also fix any chosen collar, thereby having disjoint support from the collar twist.

1.2. Collar twists of products and connected sums

The following proposition shows that collar twisting maps behave well with respect to products and connected sums. Here and in what follows, we identify (MN)(M\sharp N)^{\circ} with the boundary connected sum MNM^{\circ}\natural N^{\circ} via the preferred isotopy class of diffeomorphisms between these two manifolds.

Proposition 1.1.

Let MM and NN be closed oriented connected manifolds of dimension mm and nn.

  1. (i)

    The compositions

    SO(m)SO(m+n)ΥM×NBDiff((M×N))\mathrm{SO}(m)\subset\mathrm{SO}(m+n)\overset{\Upsilon_{M\times N}}{\longrightarrow}\mathrm{BDiff}_{\partial}((M\times N)^{\circ})

    and

    SO(m)ΥMBDiff(M)()×idNBDiff(M×N)extBDiff((M×N))\mathrm{SO}(m)\overset{\Upsilon_{M}}{\longrightarrow}\mathrm{BDiff}_{\partial}(M^{\circ})\xrightarrow{(-)\times\mathrm{id}_{N}}\mathrm{BDiff}_{\partial}(M^{\circ}\times N)\overset{\mathrm{ext}}{\longrightarrow}\mathrm{BDiff}_{\partial}((M\times N)^{\circ})

    are homotopic. In particular,

    tM×N=tM×idNπ0Diff((M×N))for m2.t_{M\times N}=t_{M}\times\mathrm{id}_{N}\in\pi_{0}\,\mathrm{Diff}_{\partial}((M\times N)^{\circ})\quad\text{for }m\geq 2.
  2. (ii)

    If MM and NN are of the same dimension d=m=nd=m=n, then the map

    SO(d)ΥMNBDiff((MN))=BDiff(MN)\mathrm{SO}(d)\overset{\Upsilon_{M\sharp N}}{\longrightarrow}\mathrm{BDiff}_{\partial}((M\sharp N)^{\circ})=\mathrm{BDiff}_{\partial}(M^{\circ}\natural N^{\circ})

    and the composition

    SO(d)diagSO(d)×SO(d)ΥM×ΥNBDiff(M)×BDiff(N)()()BDiff(MN)\hskip 34.14322pt\mathrm{SO}(d)\overset{{\mathrm{diag}}}{\longrightarrow}\mathrm{SO}(d)\times\mathrm{SO}(d)\xrightarrow{\Upsilon_{M}\times\Upsilon_{N}}\mathrm{BDiff}_{\partial}(M^{\circ})\times\mathrm{BDiff}_{\partial}(N^{\circ})\xrightarrow{(-)\natural(-)}\mathrm{BDiff}_{\partial}(M^{\circ}\natural N^{\circ})

    are homotopic after restriction to the subspace SO(d1)SO(d)\mathrm{SO}(d-1)\subset\mathrm{SO}(d). In particular, we have

    tMN=(tMidN)+(idMtN)π0Diff(MN)for d3.t_{M\sharp N}=(t_{M}\natural\mathrm{id}_{N^{\circ}})+(\mathrm{id}_{M^{\circ}}\natural t_{N})\in\pi_{0}\,\mathrm{Diff}_{\partial}(M^{\circ}\natural N^{\circ})\quad\text{for }d\geq 3.

In order to prove 1.1, it is convenient to view the collar twisting map as the instance P=P=\ast of a more general construction for a compact smooth pp-dimensional manifold PP equipped with an embedding P×DdpMP\times D^{d-p}\subset M. First, one extends the latter inclusion to an embedding P¯×D2dpM\smash{\overline{P}\times D_{2}^{d-p}\subset M} where D2dp𝐑dp\smash{D_{2}^{d-p}\subset\mathbf{R}^{d-p}} is the disc of radius 22 and P¯PP×{0}(P×[0,1])\overline{P}\coloneqq P\cup_{\partial P\times\{0\}}(\partial P\times[0,1]) is obtained by attaching an external collar to PP. This extension is unique up to isotopy. Given a smooth function λ:P¯×[0,2][0,1]\lambda\colon\smash{\overline{P}}\times[0,2]\to[0,1] and a smooth loop γΩSO(dp)\gamma\in\Omega\mathrm{SO}(d-p) that is constant near the endpoints, consider the diffeomorphism ϕλ(γ):P¯×D2dpP¯×D2dp\smash{\phi_{\lambda}(\gamma)\colon\overline{P}\times D_{2}^{d-p}\rightarrow\overline{P}\times D_{2}^{d-p}} by sending (p,x)(p,x) to (p,γ(λ(p,x))x)(p,\gamma(\lambda(p,\|x\|))\cdot x). In other words, thinking of P¯×D2dp\smash{\overline{P}\times D_{2}^{d-p}} as foliated by the leaves Sp,r:-{p}×Drdp\smash{S_{p,r}\coloneq\{p\}\times D_{r}^{d-p}} for pP¯\smash{p\in\overline{P}} and r[0,2]r\in[0,2], the diffeomorphism ϕλ(γ)\phi_{\lambda}(\gamma) preserves the leaves and acts on the leaf Sp,rS_{p,r} by rotation with the element at time λ(p,r)\lambda(p,r) of the loop γ\gamma. If one additionally assumes that

  1. (i)

    λ=1\lambda=1 on a neighbourhood of P×DdpP\times D^{d-p} where Ddp=D1dpD2dp\smash{D^{d-p}=D^{d-p}_{1}\subset D_{2}^{d-p}} is the unit disc,

  2. (ii)

    λ=0\lambda=0 on a neighbourhood of (P¯×D2dp)\partial(\overline{P}\times D_{2}^{d-p}),

then ϕλ(γ)\phi_{\lambda}(\gamma) agrees with the identity on a neighbourhood of P×DdpP¯×D2dpP\times D^{d-p}\subset\overline{P}\times D_{2}^{d-p} so restricts to a diffeomorphism of the complement. This diffeomorphism of the complement extends via the identity to a diffeomorphism of M\int(P×Ddp)M\backslash{\mathrm{int}}(P\times D^{d-p}) fixing a neighbourhood of the boundary pointwise, so we obtain a map

ϕλ():ΩSO(dp)Diff(M\int(P×Ddp))\phi_{\lambda}(-)\colon\Omega\mathrm{SO}(d-p)\longrightarrow\mathrm{Diff}_{\partial}(M\backslash{\mathrm{int}}(P\times D^{d-p}))

which depends continuously on λ\lambda and is a homomorphism with respect to pointwise multiplication on the domain and composition on the target. Since the space of smooth functions λ\lambda satisfying (i) and (ii) is contractible by linear interpolation, the delooping of ϕλ()\phi_{\lambda}(-)

ΦP:SO(dp)BDiff(M\int(P×Ddp))\Phi_{P}\colon\mathrm{SO}(d-p)\longrightarrow\mathrm{BDiff}_{\partial}(M\backslash{\mathrm{int}}(P\times D^{d-p}))

is independent of λ\lambda up to homotopy, so only depends on the isotopy class of the embedding P×DdpMP\times D^{d-p}\subset M. This map generalises the collar twisting map in the following sense.

Lemma 1.2.

For 0pd0\leq p\leq d, the maps

ΥM|SO(dp):SO(dp)BDiff(M)andΦDp:SO(dp)BDiff(M\int(Dp×Ddp))extBDiff(M).\begin{array}[]{c@{\hskip 0.05cm} l@{\hskip 0.05cm} c@{\hskip 0.1cm} l@{\hskip 0.1cm} l@{\hskip 0cm} l@{\hskip 0cm}}\Upsilon_{M}|_{\mathrm{SO}(d-p)}\hfil\hskip 1.42271pt&\colon\hfil\hskip 1.42271pt&\mathrm{SO}(d-p)\hfil\hskip 2.84544pt&\longrightarrow\hfil\hskip 2.84544pt&\mathrm{BDiff}_{\partial}(M^{\circ})\quad\text{and}\hfil\hskip 0.0pt\\ \Phi_{D^{p}}\hfil\hskip 1.42271pt&\colon\hfil\hskip 1.42271pt&\mathrm{SO}(d-p)\hfil\hskip 2.84544pt&\longrightarrow\hfil\hskip 2.84544pt&\mathrm{BDiff}_{\partial}(M\backslash{\mathrm{int}}(D^{p}\times D^{d-p}))\overset{\mathrm{ext}}{\simeq}\mathrm{BDiff}_{\partial}(M^{\circ}).\hfil\hskip 0.0pt\\ \end{array}

are homotopic. Here the embedding Dp×DdpMD^{p}\times D^{d-p}\subset M is chosen to be compatible with the orientation.

Proof.

It suffices to show that the two maps ΩSO(dp)Diff(M)\Omega\mathrm{SO}(d-p)\rightarrow\mathrm{Diff}_{\partial}(M^{\circ}) before delooping are homotopic as maps of topological groups. Going through the construction, one sees that both maps are instances of the following construction applied to smooth loops γSO(dp)\gamma\in\mathrm{SO}(d-p) that are constant near the ends: pick a smooth map λ:D2p×[0,2][0,1]\smash{\lambda\colon D^{p}_{2}\times[0,2]\to[0,1]} which is 0 in a neighbourhood of (D2p×D2dp)\smash{\partial(D^{p}_{2}\times D_{2}^{d-p})}, and 11 in a neighbourhood of DdD^{d}, consider the self-diffeomorphism of D2p2×D2p\smash{D^{p-2}_{2}\times D_{2}^{p}} sending (p,x)(p,x) to (p,γ(λ(p,x))x)(p,\gamma(\lambda(p,\|x\|))\cdot x), restrict it to a diffeomorphism of (D2p2×Dp)\int(Dd)\smash{(D^{p-2}_{2}\times D^{p})\backslash\mathrm{int}(D^{d})}, and extend the result to a diffeomorphism of MM^{\circ} by the identity. As the space of choices for λ\lambda is contractible by linear interpolation, all maps constructed this way are homotopic. ∎

A similar argument also shows the following naturality property of the map ΦP\Phi_{P}.

Lemma 1.3.

Given a compact submanifold Qint(P)Q\subset\mathrm{int}(P) of codimension 0, the map

SO(dp)ΦQBDiff(M\int(Q×Ddp))\mathrm{SO}(d-p)\overset{\Phi_{Q}}{\longrightarrow}\mathrm{BDiff}_{\partial}(M\backslash{\mathrm{int}}(Q\times D^{d-p}))

and the composition

SO(dp)ΦPBDiff(M\int(P×Ddp))extBDiff(M\int(Q×Ddp))\mathrm{SO}(d-p)\overset{\Phi_{P}}{\longrightarrow}\mathrm{BDiff}_{\partial}(M\backslash\mathrm{int}(P\times D^{d-p}))\overset{\mathrm{ext}}{\longrightarrow}\mathrm{BDiff}_{\partial}(M\backslash\mathrm{int}(Q\times D^{d-p}))

are homotopic. Here the embedding Q×DdpMQ\times D^{d-p}\subset M is the restriction of the embedding P×DdpMP\times D^{d-p}\subset M.

Equipped with Lemmas 1.2 and 1.3, we now turn to the proof of 1.1.

Proof of 1.1.

For part (i), note that the composition SO(m)BDiff(M×N)\mathrm{SO}(m)\to\mathrm{BDiff}_{\partial}(M^{\circ}\times N) is an instance of ΦN\Phi_{N} using the embedding Dm×NM×ND^{m}\times N\subset M\times N, so its postcomposition with ext:BDiff(M×N)BDiff((M×N)\int(Dm×Dn))\mathrm{ext}\colon\mathrm{BDiff}_{\partial}(M^{\circ}\times N)\rightarrow\mathrm{BDiff}_{\partial}((M\times N)\backslash\mathrm{int}(D^{m}\times D^{n})) is homotopic to ΦDn\Phi_{D^{n}} by 1.3, which in turn implies the claim as a result of 1.2. For part (ii), view (MN)(M\sharp N)^{\circ} as being obtained from MNM^{\circ}\sqcup N^{\circ} by gluing on a pair-of-pants bordism W:Sd1Sd1Sd1W\colon S^{d-1}\sqcup S^{d-1}\leadsto S^{d-1}. To show the claim, it suffices to show that the maps tin,tout:SO(d1)BDiff(W)t_{\mathrm{in}},t_{\mathrm{out}}\colon\mathrm{SO}(d-1)\rightarrow\mathrm{BDiff}_{\partial}(W) are homotopic, where tint_{\mathrm{in}} simultaneously twists collars of the two incoming boundary spheres and toutt_{\mathrm{out}} twists a collar of the outgoing boundary sphere. Viewing WW as Dd\int((e(D1D1))×Dd1)D^{d}\backslash\mathrm{int}((e(D^{1}\sqcup D^{1}))\times D^{d-1}) for an embedding e:D1D1int(D1)e\colon D^{1}\sqcup D^{1}\hookrightarrow\mathrm{int}(D^{1}), the map tint_{\mathrm{in}} is given by ΥD1D1:SO(d1)BDiff(Dd\int(e(D1D1)×Dd1))\Upsilon_{D^{1}\sqcup D^{1}}\colon\mathrm{SO}(d-1)\rightarrow\mathrm{BDiff}_{\partial}(D^{d}\backslash\mathrm{int}(e(D^{1}\sqcup D^{1})\times D^{d-1})) and the map toutt_{{\mathrm{out}}} as the composition of ΥD1:SO(d1)BDiff(Dd\int(D1×Dd1)\Upsilon_{D^{1}}\colon\mathrm{SO}(d-1)\rightarrow\mathrm{BDiff}_{\partial}(D^{d}\backslash\mathrm{int}(D^{1}\times D^{d-1}) with ext:BDiff(Dd\int(D1×Dd1))BDiff(Dd\int((e(D1D1))×Dd1))\mathrm{ext}\colon\mathrm{BDiff}_{\partial}(D^{d}\backslash\mathrm{int}(D^{1}\times D^{d-1}))\rightarrow\mathrm{BDiff}_{\partial}(D^{d}\backslash\mathrm{int}((e(D^{1}\sqcup D^{1}))\times D^{d-1})), so the claim follows from 1.3 applied to P=D1P=D^{1}, Q=D1D1Q=D^{1}\sqcup D^{1}, and M=DdM=D^{d}. ∎

Remark 1.4.

1.1 (ii) is a more general version of the “pants relation" in [CT22, Lemma 2.5], where a proof of this relation is given by constructing an explicit isotopy.

1.3. Collar twists of exotic spheres

We now turn to the collar twisting map ΥΣ\Upsilon_{\Sigma} for homotopy spheres Σ\Sigma, but we actually restrict our attention to the collar twist tΣπ0Diff(Σ)t_{\Sigma}\in\pi_{0}\,\mathrm{Diff}_{\partial}(\Sigma^{\circ}) it induces on fundamental groups. We begin with a recollection of the classification of homotopy spheres.

1.3.1. Classification of homotopy spheres

Recall (e.g. from [Lev85, p. 90-91]) that Kervaire–Milnor’s finite abelian group Θd\Theta_{d} of homotopy dd-spheres [KM63] fits for d5d\geq 5 into an exact sequence

(4) 0bPd+1Θd[]coker(J)d{𝐙/2if d=2k2, for some k0otherwise0\rightarrow\mathrm{bP}_{d+1}\rightarrow\Theta_{d}\xrightarrow{[-]}\mathrm{coker}(J)_{d}\rightarrow\begin{cases}\mathbf{Z}/2&\text{if }d=2^{k}-2,\text{ for some }k\\ 0&\text{otherwise}\end{cases}

where bPd+1Θd\mathrm{bP}_{d+1}\leq\Theta_{d} is a certain cyclic subgroup and coker(J)d\mathrm{coker}(J)_{d} is the cokernel of the stable JJ-homomorphism πdOπd𝐒\pi_{d}\,\mathrm{O}\rightarrow\pi_{d}\,\mathbf{S} from the homotopy groups of the stable orthogonal group (which are known by Bott periodicity) to the stable homotopy groups of spheres. The order of the cyclic subgroup bPd+1\mathrm{bP}_{d+1} is known in all cases except d=125d=125 (combine [Lev85, Corollaries 2.2, 3.20, Theorem 4.9] with [HHR16, Theorem 1.3]):

bPd+1={22k2(22k11)num(4|B2k|/k)if d=4k1 for k22if d=4k+1 for k1 but d2k3 if k70if d is even or if d=2k3 for k62 or 0if d=2k3 for k=7.\sharp\mathrm{bP}_{d+1}=\begin{cases}2^{2k-2}(2^{2k-1}-1)\,\mathrm{num}(4|B_{2k}|/k)&\text{if }d=4k-1\text{ for $k\geq 2$}\\ 2&\text{if }d=4k+1\text{ for $k\geq 1$}\text{ but }d\neq 2^{k}-3\text{ if }k\leq 7\\ 0&\text{if }d\text{ is even or if }d=2^{k}-3\text{ for }k\leq 6\\ 2\text{ or }0&\text{if }d=2^{k}-3\text{ for }k=7.\end{cases}

The map coker(J)d𝐙/2\mathrm{coker}(J)_{d}\rightarrow\mathbf{Z}/2 in the sequence (4) is known to be trivial as long as d2k2d\neq 2^{k}-2 for k>7k>7. It is known to be nontrivial for k6k\leq 6, but the case k=7k=7 (i.e. d=126d=126) is still open (see [HHR16, Theorem 1.4]). The question whether bP126=0\mathrm{bP}_{126}=0 or bP126=𝐙/2\mathrm{bP}_{126}=\mathbf{Z}/2 and the question whether coker(J)126𝐙/2\mathrm{coker}(J)_{126}\rightarrow\mathbf{Z}/2 is surjective or not (these questions turn out to be equivalent; see [Lev70, p. 88]) is the last remaining case of the Kervaire invariant one problem. The upshot of this discussion is that apart from the two problematic dimensions d=125,126d=125,126, the group Θd\Theta_{d} is described in terms of the group coker(J)d\mathrm{coker}(J)_{d} up to extension problems. In most cases, also these extension problems have been resolved:

  • \bullet

    For dd even, bPd+1\mathrm{bP}_{d+1} vanishes and the map to Θd[]coker(J)d\Theta_{d}\xrightarrow{[-]}\mathrm{coker}(J)_{d} is an isomorphism as long as d2k2d\neq 2^{k}-2 for k>7k>7, so in these cases there are no extension problems.

  • \bullet

    For d=2k2d=2^{k}-2 with k6k\leq 6, we have an exact sequence 0Θdcoker(J)d𝐙/200\rightarrow\Theta_{d}\rightarrow\mathrm{coker}(J)_{d}\rightarrow\mathbf{Z}/2\rightarrow 0 which admits a splitting since in these dimensions coker(J)d\mathrm{coker}(J)_{d} is known to be annihilated by 22 (see e.g. the table [IWX20, Table 1]), so Θdcoker(J)d𝐙/2\Theta_{d}\cong\mathrm{coker}(J)_{d}\oplus\mathbf{Z}/2. For k=7k=7 the question whether the map coker(J)126𝐙/2\mathrm{coker}(J)_{126}\rightarrow\mathbf{Z}/2 is split surjective (rather than just surjective which is open too; see above) is known as the strong Kervaire invariant one problem.

  • \bullet

    For d3(mod4)d\equiv 3\pmod{4}, the map Θdcoker(J)d\Theta_{d}\rightarrow\mathrm{coker}(J)_{d} is split surjective by [Bru68, Theorem 1.3] or [Fra73, Theorem 5], so ΘdbPd+1coker(J)d\Theta_{d}\cong\mathrm{bP}_{d+1}\oplus\mathrm{coker}(J)_{d}.

  • \bullet

    For d1(mod4)d\equiv 1\pmod{4}, the map Θdcoker(J)d\Theta_{d}\rightarrow\mathrm{coker}(J)_{d} is split surjective if dd is not of the form 2k32^{k}-3 for some k1k\geq 1 by [Bru69, Theorem 1.2] and [Bru70, Theorem 1.1], so Θd𝐙/2coker(J)d\Theta_{d}\cong\mathbf{Z}/2\oplus\mathrm{coker}(J)_{d}.

1.3.2. Collar twists of homotopy spheres and the Milnor–Munkres–Novikov pairing

We begin the discussion of collar twists of homotopy spheres with a general observation: if MM is a closed oriented manifold of dimension d5d\geq 5 and ΣΘd\Sigma\in\Theta_{d} is a homotopy sphere, then writing Σ¯Θd\overline{\Sigma}\in\Theta_{d} for the inverse sphere obtained by reversing the orientation, the maps

()idΣ:BDiff(M)BDiff(MΣ)=BDiff((MΣ))()idΣ¯:BDiff((MΣ))BDiff((MΣ)Σ¯)=BDiff(M).\begin{array}[]{c@{\hskip 0.05cm} l@{\hskip 0.05cm} c@{\hskip 0.1cm} l@{\hskip 0.1cm} l@{\hskip 0cm} l@{\hskip 0cm}}(-)\natural\mathrm{id}_{\Sigma^{\circ}}\hfil\hskip 1.42271pt&\colon\hfil\hskip 1.42271pt&\mathrm{BDiff}_{\partial}(M^{\circ})\hfil\hskip 2.84544pt&\longrightarrow\hfil\hskip 2.84544pt&\mathrm{BDiff}_{\partial}(M^{\circ}\natural\Sigma^{\circ})=\mathrm{BDiff}_{\partial}((M\sharp\Sigma)^{\circ})\hfil\hskip 0.0pt\\ (-)\natural\mathrm{id}_{\overline{\Sigma}^{\circ}}\hfil\hskip 1.42271pt&\colon\hfil\hskip 1.42271pt&\mathrm{BDiff}_{\partial}((M\sharp\Sigma)^{\circ})\hfil\hskip 2.84544pt&\longrightarrow\hfil\hskip 2.84544pt&\mathrm{BDiff}_{\partial}((M\sharp\Sigma)^{\circ}\natural\overline{\Sigma}^{\circ})=\mathrm{BDiff}_{\partial}(M^{\circ}).\hfil\hskip 0.0pt\\ \end{array}

are inverse homotopy equivalences, so in particular induce an isomorphism π0Diff(M)π0Diff((MΣ))\pi_{0}\,\mathrm{Diff}_{\partial}(M^{\circ})\cong\pi_{0}\,\mathrm{Diff}_{\partial}((M\sharp\Sigma)^{\circ}) on fundamental groups. For M=SdM=S^{d}, combining the latter with the usual isomorphism π0Diff(Dd)Θd+1\pi_{0}\,\mathrm{Diff}_{\partial}(D^{d})\cong\Theta_{d+1} given by gluing together two copies Dd+1D^{d+1} along their boundary via diffeomorphisms of SdS^{d} supported on a hemisphere results in a chain of isomorphisms

π0Diff(Σ)π0Diff(Dd)Θd+1\pi_{0}\,\mathrm{Diff}_{\partial}(\Sigma^{\circ})\cong\pi_{0}\,\mathrm{Diff}_{\partial}(D^{d})\cong\Theta_{d+1}

We write

TΣΘd+1T_{\Sigma}\in\Theta_{d+1}

for the image of the collar twist tΣπ0Diff(Σ)t_{\Sigma}\in\pi_{0}\,\mathrm{Diff}_{\partial}(\Sigma^{\circ}) under these isomorphisms. This defines a set-theoretical function T():ΘdΘd+1T_{(-)}\colon\Theta_{d}\rightarrow\Theta_{d+1} which can be rephrased (see 1.5 below) in terms of a well-known construction in the study of homotopy spheres, namely the bilinear Milnor–Munkres–Novikov pairing (see e.g. [Bre67]) πk𝐒ΘdΘk+d\pi_{k}\,\mathbf{S}\otimes\Theta_{d}\rightarrow\Theta_{k+d} for k<d1k<d-1. The latter is related to the multiplication in the stable homotopy groups of spheres by a commutative diagram

(5) πk𝐒ΘdΘk+dπk𝐒coker(J)dcoker(J)k+d()()idπk𝐒[][]()() for k<d1\leavevmode\hbox to172.47pt{\vbox to51.98pt{\pgfpicture\makeatletter\hbox{\hskip 86.23569pt\lower-26.04149pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{}{}{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{\offinterlineskip{}{}{{{}}{{}}{{}}{{}}}{{{}}}{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-86.23569pt}{-25.94165pt}\pgfsys@invoke{ }\hbox{\vbox{\halign{\pgf@matrix@init@row\pgf@matrix@step@column{\pgf@matrix@startcell#\pgf@matrix@endcell}&#\pgf@matrix@padding&&\pgf@matrix@step@column{\pgf@matrix@startcell#\pgf@matrix@endcell}&#\pgf@matrix@padding\cr\hfil\hskip 22.65884pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-18.3533pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${\pi_{k}\,\mathbf{S}\otimes\Theta_{d}}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}}}&\hskip 22.65884pt\hfil&\hfil\hskip 57.37538pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-9.06987pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${\Theta_{k+d}}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}&\hskip 13.37541pt\hfil\cr\vskip 18.00005pt\cr\hfil\hskip 36.20403pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-31.89848pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${\pi_{k}\,\mathbf{S}\otimes\mathrm{coker}(J)_{d}}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}&\hskip 36.20403pt\hfil&\hfil\hskip 72.03165pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-23.72614pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${\mathrm{coker}(J)_{k+d}}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}&\hskip 28.03168pt\hfil\cr}}}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}}{{{{}}}{{}}{{}}{{}}{{}}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} {}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{{ {\pgfsys@beginscope \pgfsys@setdash{}{0.0pt}\pgfsys@roundcap\pgfsys@roundjoin{} {}{}{} {}{}{} \pgfsys@moveto{-2.07988pt}{2.39986pt}\pgfsys@curveto{-1.69989pt}{0.95992pt}{-0.85313pt}{0.27998pt}{0.0pt}{0.0pt}\pgfsys@curveto{-0.85313pt}{-0.27998pt}{-1.69989pt}{-0.95992pt}{-2.07988pt}{-2.39986pt}\pgfsys@stroke\pgfsys@endscope}} }{}{}{{}}{}{}{{}}\pgfsys@moveto{-27.17282pt}{11.78893pt}\pgfsys@lineto{44.22864pt}{11.78893pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{44.42862pt}{11.78893pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-0.02211pt}{15.8917pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$\scriptstyle{(-)\cdot(-)}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{-50.03166pt}{3.48479pt}\pgfsys@lineto{-50.03166pt}{-14.18198pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{0.0}{-1.0}{1.0}{0.0}{-50.03166pt}{-14.38196pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{{}{}}}{{}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-76.15344pt}{-6.9347pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$\scriptstyle{\mathrm{id}_{\pi_{k}\mathbf{S}}\otimes[-]}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{58.20401pt}{3.01811pt}\pgfsys@lineto{58.20401pt}{-14.18198pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{0.0}{-1.0}{1.0}{0.0}{58.20401pt}{-14.38196pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{60.55678pt}{-7.5319pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$\scriptstyle{[-]}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{-13.62764pt}{-23.44165pt}\pgfsys@lineto{29.57237pt}{-23.44165pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{29.77235pt}{-23.44165pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-0.57765pt}{-19.33888pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$\scriptstyle{(-)\cdot(-)}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}\quad\text{ for }k<d-1

with bottom horizontal map induced by the multiplication on the stable stems, using that products of elements in im(J)k\mathrm{im}(J)_{k} and πd𝐒\pi_{d}\,\mathbf{S} contained in im(J)k+d\mathrm{im}(J)_{k+d} if k<d1k<d-1 (see p. 442 of loc.cit.).

Proposition 1.5 (Kreck, Levine).

We have TΣ=ηΣT_{\Sigma}=\eta\cdot\Sigma where ηπ1𝐒𝐙/2\eta\in\pi_{1}\,\mathbf{S}\cong\mathbf{Z}/2 is the generator.

Proof.

Levine writes γ(Σ)Θd+1\gamma(\Sigma)\in\Theta_{d+1} for TΣΘd+1T_{\Sigma}\in\Theta_{d+1} [Lev70, p. 245-246] and Kreck writes ΣΣΘd+1\Sigma_{\Sigma}\in\Theta_{d+1} for it [Kre79, p. 646]. For even dd, the claim is [Kre79, Lemma 3 c)]. For odd dd, the subgroup bPd+2Θd+1\mathrm{bP}_{d+2}\leq\Theta_{d+1} is trivial, so it suffices to show the claimed equality after passing to coker(J)d+1\mathrm{coker}(J)_{d+1} (see Section 1.3.1). The latter follows from [Lev70, Corollary 4] using that Levine’s subgroup I1(Σ)Θd+1I_{1}(\Sigma)\subset\Theta_{d+1} is generated by γ(Σ)Θd+1\gamma(\Sigma)\in\Theta_{d+1} by definition; see p. 246 loc.cit.. ∎

Remark 1.6.

1.5 has immediate consequences for collar twists of homotopy spheres. For example, since η\eta is 22-torsion and the Milnor–Munkres–Novikov pairing is bilinear, the sphere TΣ=ηΣT_{\Sigma}=\eta\cdot\Sigma is trivial if ΣΘd\Sigma\in\Theta_{d} has odd order, so the collar twist of Σ\Sigma is in these cases trivial too.

The combination of 1.5, the classification of homotopy spheres as recalled in Section 1.3.1, and the diagram (5) allows one to reduce most questions on collar twists of exotic spheres to questions in stable homotopy theory. As an example of this principle, we rephrase the condition featuring in the statements of Theorem A and D (whether TΣ=ηΣΘd+1T_{\Sigma}=\eta\cdot\Sigma\in\Theta_{d+1} is divisible by 22 or not) in most cases in terms of the cokernel of the stable JJ-homomorphism:

Lemma 1.7.

If ηΣΘd+1\eta\cdot\Sigma\in\Theta_{d+1} is divisible by 22, then so is η[Σ]coker(J)d+1\eta\cdot[\Sigma]\in\mathrm{coker}(J)_{d+1}. The converse holds

  1. (i)

    for d4,5(mod8)d\not\equiv 4,5\pmod{8},

  2. (ii)

    for d5(mod8)d\equiv 5\ \ \ \,\pmod{8} for d125d\neq 125, and

  3. (iii)

    for d4(mod8)d\equiv 4\ \ \ \,\pmod{8} for d=2k4d=2^{k}-4 with k6k\leq 6.

Proof.

By commutativity of (5), the class η[Σ]coker(J)d+1\eta\cdot[\Sigma]\in\mathrm{coker}(J)_{d+1} is the image of ηΣΘd+1\eta\cdot\Sigma\in\Theta_{d+1} under the morphism []:Θd+1coker(J)d[-]\colon\Theta_{d+1}\rightarrow\mathrm{coker}(J)_{d}, so if the latter is divisible by 22, then so is the former. To prove the partial converse, we distinguish some cases and make frequent use of the classification of homotopy spheres as recalled in Section 1.3.1, without further reference.

  • \bullet

    For d+13,7mod 8d+1\equiv 3,7\ \mathrm{mod}\ 8, the map []:Θd+1coker(J)d+1[-]\colon\Theta_{d+1}\rightarrow\mathrm{coker}(J)_{d+1} is split surjective, so Θd+1bPd+2coker(J)d+1\Theta_{d+1}\cong\mathrm{bP}_{d+2}\oplus\mathrm{coker}(J)_{d+1}. Since ηΣ\eta\cdot\Sigma has order two and bPd+2\mathrm{bP}_{d+2} is cyclic of order divisible by 44, the bPd+2\mathrm{bP}_{d+2}-component of the order 22 element ηΣ\eta\cdot\Sigma has to be divisible by 22, so the full element ηΣ\eta\cdot\Sigma is divisible by 22 if and only if its image η[Σ]coker(J)d+1\eta\cdot[\Sigma]\in\mathrm{coker}(J)_{d+1} is divisible by 22.

  • \bullet

    For d+11mod 8d+1\equiv 1\ \mathrm{mod}\ 8, the map []:Θd+1coker(J)d+1[-]\colon\Theta_{d+1}\rightarrow\mathrm{coker}(J)_{d+1} is also split surjective, so Θd+1bPd+2coker(J)d+1\Theta_{d+1}\cong\mathrm{bP}_{d+2}\oplus\mathrm{coker}(J)_{d+1} . In this case the bPd+2\mathrm{bP}_{d+2}-component of ηΣΘd+1\eta\cdot\Sigma\in\Theta_{d+1} turns out to vanish, which implies the result. The reason for this vanishing is that ηΣΘd+1\eta\cdot\Sigma\in\Theta_{d+1} is contained in the subgroup bSpind+2Θd+1\mathrm{bSpin}_{d+2}\leq\Theta_{d+1} of homotopy spheres that bound a spin manifold [Law73, §4 + Diagram (6)] and on this subgroup the bPd+2\mathrm{bP}_{d+2}-component with respect to the can be computed as the image of the ff-invariant from [Bru69, §3] which vanishes for ηΣ\eta\cdot\Sigma by [Law73, Proposition 4.1] (this uses that the pairings denoted τn,k\tau_{n,k} and ρn,k\rho_{n,k} in loc.cit.  are compatible, by diagram (B) on p. 835 of loc.cit.).

  • \bullet

    For d+10,2,4,6mod 8d+1\equiv 0,2,4,6\ \mathrm{mod}\ 8 and d+12k2d+1\neq 2^{k}-2 for k7k\leq 7, and for d+15(mod8)d+1\equiv 5\pmod{8} with d+1=2k3d+1=2^{k}-3 for k6k\leq 6 we have Θd+1coker(J)d+1\Theta_{d+1}\cong\mathrm{coker}(J)_{d+1} and there is nothing to show.

  • \bullet

    For d+16(mod8)d+1\equiv 6\pmod{8} with d+1=2k2d+1=2^{k}-2 for k6k\leq 6 we have Θd+1coker(J)d+1𝐙/2\Theta_{d+1}\cong\mathrm{coker}(J)_{d+1}\oplus\mathbf{Z}/2, so an element in Θd+1\Theta_{d+1} is divisible by 22 if and only if this holds for its image in coker(J)d+1\mathrm{coker}(J)_{d+1}. ∎

Remark 1.8.

To extend 1.7 to d+15(mod8)d+1\equiv 5\pmod{8} for d+12k3d+1\neq 2^{k}-3, it would suffice to show that the bPd+2\mathrm{bP}_{d+2}-component of ηΣ\eta\cdot\Sigma for ΣΘd\Sigma\in\Theta_{d} under the splitting Θd+1coker(J)d+1bPd+2\Theta_{d+1}\cong\mathrm{coker}(J)_{d+1}\oplus\mathrm{bP}_{d+2} recalled in Section 1.3.1 is trivial. We do not know whether this is the case.

In view of 1.7, the question whether ηΣΘd+1\eta\cdot\Sigma\in\Theta_{d+1} is divisible by 22 can in many dimensions be analysed with inputs from stable homotopy theory. The following two remarks contain some applications in this direction:

Remark 1.9.

As η\eta has order two, whether ηΣΘd+1\eta\cdot\Sigma\in\Theta_{d+1} is divisible by 22 or not can be tested 22-locally. At the prime 22, the groups coker(J)d\mathrm{coker}(J)_{d} and multiplication by η\eta on them have been computed up to dimensions about 9090. The result is summarised in [IWX20, Figure 1] where every dot represents a nontrivial element, the diagonal and vertical lines indicate that two elements are related by multiplication with η\eta or 22, respectively, and the image of JJ consists of the blue dots, apart from the blue dots in degrees 1,2(mod8)\equiv 1,2\pmod{8}. Combining this with 1.7 and the classification of homotopy spheres recalled in Section 1.3.1, one can in most dimensions up to about 9090 determine the groups Θd\Theta_{d} and the subgroups ΘdsplitΘd\smash{\Theta^{\mathrm{split}}_{d}\leq\Theta_{d}} of those ΣΘd\Sigma\in\Theta_{d} such that ηΣ\eta\cdot\Sigma is divisible by 22. The result of this analysis for d19d\leq 19 is recorded in Table 1 of the introduction.

Remark 1.10.

There are also many infinite families of homotopy spheres ΣΘd\Sigma\in\Theta_{d} for which one can decide whether ηΣΘd+1\eta\cdot\Sigma\in\Theta_{d+1} is divisible by 22 or not. We again rely on Section 1.3.1.

  • \bullet

    As an infinite family of nontrivial ΣΘd\Sigma\in\Theta_{d} in odd dimensions such that ηΣ\eta\cdot\Sigma is divisible by 22, one may for instance take any ΣΘd\Sigma\in\Theta_{d} for d1,3,7(mod8)d\equiv 1,3,7\pmod{8} that lies in the nontrivial subgroup bPd+1Θd\mathrm{bP}_{d+1}\leq\Theta_{d}. This is because ηΣΘd+1coker(J)d+1\eta\cdot\Sigma\in\Theta_{d+1}\cong\mathrm{coker}(J)_{d+1} is trivial as a result of (5), so it is in particular divisible by 22. There are also examples in even dimensions: as bP8k+3=0\mathrm{bP}_{8k+3}=0, the class in coker(J)8k+2\mathrm{coker}(J)_{8k+2} of Adams’ element μ8k+2π8k+2𝐒\mu_{8k+2}\in\pi_{8k+2}\,\mathbf{S} (which is nontrivial in coker(J)8k+2\mathrm{coker}(J)_{8k+2} as π8k+2O=0\pi_{8k+2}\,\mathrm{O}=0) lifts uniquely to a homotopy sphere Σμ8k+2Θ8k+2\Sigma_{\mu_{8k+2}}\in\Theta_{8k+2}. As η[μ8k+2]=0coker(J)8k+2\eta\cdot[\mu_{8k+2}]=0\in\mathrm{coker}(J)_{8k+2} since ημ8k+2π8k+3𝐒\eta\cdot\mu_{8k+2}\in\pi_{8k+3}\,\mathbf{S} is known to be contained in im(J)8k+3\mathrm{im}(J)_{8k+3}, it follows from 1.7 that Σμ8k+2\Sigma_{\mu_{8k+2}} is divisible by 22.

  • \bullet

    As an infinite family of nontrivial ΣΘd\Sigma\in\Theta_{d} in odd dimensions such that ηΣ\eta\cdot\Sigma is not divisible by 22, one may take any ΣΘ8k+1\Sigma\in\Theta_{8k+1} that maps to the class in coker(J)8k+1\mathrm{coker}(J)_{8k+1} represented by Adams’ element μ8k+1π8k+1𝐒\mu_{8k+1}\in\pi_{8k+1}\,\mathbf{S} which is known to have the property that ημ8k+1coker(J)8k+2=π8k+2𝐒\eta\cdot\mu_{8k+1}\in\mathrm{coker}(J)_{8k+2}=\pi_{8k+2}\,\mathbf{S} is not divisible by 22. In even dimensions, one may use the families of nontrivial homotopy spheres ΣΘd\Sigma\in\Theta_{d} for d8(mod192)d\equiv 8\pmod{192} from [Kra21, Proposition 2.11 (i)] which have the property that η[Σ]coker(J)d\eta\cdot[\Sigma]\in\mathrm{coker}(J)_{d} is nontrivial and detected in the spectrum tmf{\mathrm{tmf}} of topological modular forms (see the proof of the cited proposition). Moreover, in these dimensions πd+1tmf\pi_{d+1}\,{\mathrm{tmf}} is known to be annihilated by 22 (see e.g. [Beh20, Figure 1.2]), so η[Σ]\eta\cdot[\Sigma] is not divisible by 22 in πd+1tmf\pi_{d+1}\,{\mathrm{tmf}} and hence neither in coker(J)d+1\mathrm{coker}(J)_{d+1}.

1.4. Collar twists of tori

The next class of manifolds for which we establish some results on their collar twisting maps are homotopy tori. For this class of manifolds, it is convenient to study the fibre sequence (2) involving the collar twisting maps by comparing it to an analogous sequence for block-homeomorphisms (see e.g. [HLLRW21, Section 2] for a discussion of block-automorphisms suitable for our needs) via a map of fibre sequences

(6) BDiff(M){\mathrm{BDiff}_{\partial}(M^{\circ})}BDiff+(M){\mathrm{BDiff}^{+}_{\ast}(M)}BSO(d){\mathrm{BSO}(d)}BHomeo~TM(M){\smash{\mathrm{B\widetilde{Homeo}}}_{T_{*}M}(M)}BHomeo~+(M){\smash{\mathrm{B\widetilde{Homeo}}}^{+}_{\ast}(M)}BSTop.{\mathrm{BSTop}.}

The bottom row of this diagram deserves an explanation. To construct it, first consider the forgetful map Homeo~+(M)hAut+(M)\smash{\mathrm{\widetilde{Homeo}}}^{+}(M)\rightarrow\mathrm{hAut}^{+}(M) from the space of orientation-preserving block-homeomorphisms of MM to the space of orientation-preserving homotopy self-equivalences. The space Homeo~+(M)\smash{\mathrm{\widetilde{Homeo}}}^{+}_{\ast}(M) is defined as the homotopy pullback of this map along the inclusion map hAut+(M)hAut+(M)\mathrm{hAut}_{\ast}^{+}(M)\rightarrow\mathrm{hAut}^{+}(M) of those orientation-preserving self-equivalences of MM that preserve the chosen point M\ast\in M. The delooping of the latter map is the universal MM-fibration, so BHomeo~+(M)\smash{\mathrm{B\widetilde{Homeo}}}^{+}_{\ast}(M) is by construction equivalent to the total space of the universal oriented MM-block-bundle. The right-hand map in the bottom sequence is the delooping of the map Homeo~+(M)STop\smash{\mathrm{\widetilde{Homeo}}}^{+}_{\ast}(M)\rightarrow\mathrm{STop} that takes the stable topological derivative of a block-homeomorphism of MM at M\ast\in M, or equivalently, it classifies the stable vertical topological tangent bundle of the universal oriented MM-block-bundle (see Section 2 loc.cit.), similarly to how the right-hand map of the upper sequence classifies the vertical tangent bundle of the universal oriented smooth MM-bundle. The rightmost vertical map classifies the underlying stable Euclidean bundle of an oriented dd-dimensional vector bundle and the middle vertical map is induced by the forgetful map Diff+(M)Homeo~+(M)\mathrm{Diff}_{\ast}^{+}(M)\rightarrow\smash{\mathrm{\widetilde{Homeo}}}^{+}_{\ast}(M). The left-hand map of the bottom sequence is defined as the homotopy fibre inclusion of the right-hand map, or equivalently, as the delooping of the derivative map.

Note that the bottom row only depends on the underlying topological manifold of MM, so in particular agrees for homotopy tori M=𝒯M=\mathcal{T} with the corresponding sequence of the standard torus M=TdM=T^{d}. For the latter, the middle space BHomeo~+(M)\smash{\mathrm{B\widetilde{Homeo}}}^{+}_{\ast}(M) has a very simple description:

Lemma 1.11.

For any d1d\geq 1, the map

BHomeo~+(Td)BSLd(𝐙)\smash{\mathrm{B\widetilde{Homeo}}}^{+}_{\ast}(T^{d})\longrightarrow\mathrm{BSL}_{d}(\mathbf{Z})

induced by the action on H1(Td)𝐙d\mathrm{H}_{1}(T^{d})\cong\mathbf{Z}^{d} is an equivalence.

Proof.

As TdK(𝐙d,1)T^{d}\simeq K(\mathbf{Z}^{d},1), the analogous map hAut+(Td)SLd(𝐙)\mathrm{hAut}^{+}_{\ast}(T^{d})\rightarrow\mathrm{SL}_{d}(\mathbf{Z}) from the space of orientation homotopy self-equivalences of MM is an equivalence, so it suffices to show that the forgetful map Homeo~+(Td)hAut+(Td)\smash{\mathrm{\widetilde{Homeo}}}^{+}_{\ast}(T^{d})\rightarrow\mathrm{hAut}^{+}_{\ast}(T^{d}) is an equivalence. We will do so by proving that the right-hand map in the map of homotopy fibre sequences

Td{T^{d}}BHomeo~+(Td){\smash{\mathrm{B\widetilde{Homeo}}}^{+}_{\ast}(T^{d})}BHomeo~+(Td){\smash{\mathrm{B\widetilde{Homeo}}}^{+}(T^{d})}Td{T^{d}}BhAut+(Td){\mathrm{BhAut}^{+}_{\ast}(T^{d})}BhAut+(Td){\mathrm{BhAut}^{+}(T^{d})}

comparing the universal TdT^{d}-block-bundle with the universal TdT^{d}-fibration, is an equivalence. Using the action of SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) on Td𝐑d/𝐙dT^{d}\cong\mathbf{R}^{d}/\mathbf{Z}^{d} and the action of TdT^{d} on itself, a diagram chase in the ladder of long exact sequences induced by this map of fibre sequences shows that the middle arrow is surjective on all homotopy groups. Injectivity on homotopy groups is equivalent to the claim that for k0k\geq 0, any self-homeomorphism of Td×DkT^{d}\times D^{k} fixing on the boundary that is homotopic to the identity relative to the boundary is also concordant to the identity relative to the boundary. For d5d\geq 5, this from the fact that the topological structure sets 𝒮Top(Td×Dk)\smash{\mathcal{S}^{\mathrm{Top}}_{\partial}(T^{d}\times D^{k})} in the sense of surgery theory are trivial as long as k+d5k+d\geq 5 [KS77, p. 205, Theorem C.2], but there is also a more direct proof in all dimensions [Law76]. ∎

Corollary 1.12.

Let 𝒯\mathcal{T} be a homotopy torus of dimension d1d\geq 1. The collar twisting map

Υ𝒯:SO(d)BDiff(𝒯)\Upsilon_{\mathcal{T}}\colon\mathrm{SO}(d)\longrightarrow\mathrm{BDiff}_{\partial}(\mathcal{T}^{\circ})

is injective on πk()\pi_{k}(-) for kd2k\leq d-2. In particular, t𝒯π0Diff(𝒯)t_{\mathcal{T}}\in\pi_{0}\,\mathrm{Diff}_{\partial}(\mathcal{T}^{\circ}) is nontrivial for d3d\geq 3.

Proof.

From the map of long exact sequences induced by the map (6) together with the fact that the higher homotopy groups of BHomeo~+(𝒯)BHomeo~+(Td)\smash{\mathrm{B\widetilde{Homeo}}}^{+}_{\ast}(\mathcal{T})\simeq\smash{\mathrm{B\widetilde{Homeo}}}^{+}_{\ast}(T^{d}) vanish as a result of 1.11, we see that the map in question is injective on πk()\pi_{k}(-) if the map πkSO(d)πkTop\pi_{k}\,\mathrm{SO}(d)\rightarrow\pi_{k}\,\mathrm{Top} is injective. This maps factors as the stabilisation map SO(d)SO\mathrm{SO}(d)\rightarrow\mathrm{SO} followed by the forgetful map SOTop\mathrm{SO}\rightarrow\mathrm{Top}. The latter is injective on all homotopy groups (combine [Bru68] with [KS77, p. 246, 5.0.(1)]) and the former for kd2k\leq d-2 by stability, so the claim follows. ∎

Remark 1.13.

The collar twist t𝒯π0Diff(𝒯)t_{\mathcal{T}}\in\pi_{0}\,\mathrm{Diff}_{\partial}(\mathcal{T}^{\circ}) is also nontrivial for d=2d=2 which follows for instance from 2.2 below. Moreover, for d=2,3d=2,3 the map Υ𝒯\Upsilon_{\mathcal{T}} induces an isomorphism on all higher homotopy groups since the component of the identity Diff(Td)id\mathrm{Diff}_{*}(T^{d})_{\mathrm{id}}\simeq\ast is contractible. This is well-known for d=2d=2 and follows for d=3d=3 from a combination of [Hat83] and [Hat76] using that T3T^{3} is Haken.

Remark 1.14.

Replacing SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) with Aut(π1M)\mathrm{Aut}(\pi_{1}\,M), the statement of 1.11 (and thus also that of 1.12) holds for many other closed aspherical manifolds MM, in particular for those of dimension d5d\geq 5 whose fundamental group satisfies the Farrell–Jones conjecture and also for those of dimension d=4d=4 if the fundamental group is good in the sense of [FQ90, p. 99] (see e.g. [HLLRW21, Proposition 5.1.1] for an explanation of this).

2. Mapping class groups of exotic tori and the proof of D

Equipped with the results on collar twists from the previous section, we turn towards studying the mapping class groups of homotopy tori of the form TdΣT^{d}\sharp\Sigma.

2.1. Central extensions of special linear groups

The strategy will be to relate the mapping class groups of homotopy tori to well-known central extensions of special linear groups. We first recall these extensions and discuss some of their properties. The universal cover of the stable special linear group over the reals SL(𝐑)=colimdSLd(𝐑)\mathrm{SL}(\mathbf{R})=\mathrm{colim}_{d}\,\mathrm{SL}_{d}(\mathbf{R}) gives a central extension

0𝐙/2SL¯(𝐑)SL(𝐑)00\longrightarrow\mathbf{Z}/2\longrightarrow\overline{\mathrm{SL}}(\mathbf{R})\longrightarrow\mathrm{SL}(\mathbf{R})\longrightarrow 0

which we may pull back along the lattice inclusion SLd(𝐙)SL(𝐑)\mathrm{SL}_{d}(\mathbf{Z})\leq\mathrm{SL}(\mathbf{R}) to a central extension

(7) 0𝐙/2SL¯d(𝐙)SLd(𝐙)00\longrightarrow\mathbf{Z}/2\longrightarrow\overline{\mathrm{SL}}_{d}(\mathbf{Z})\longrightarrow\mathrm{SL}_{d}(\mathbf{Z})\longrightarrow 0

for d1d\geq 1. For d=2d=2, also a different central extension will play a role, namely the pullback

(8) 0𝐙SL~2(𝐙)SL2(𝐙)00\longrightarrow\mathbf{Z}\longrightarrow\widetilde{\mathrm{SL}}_{2}(\mathbf{Z})\longrightarrow\mathrm{SL}_{2}(\mathbf{Z})\longrightarrow 0

along the inclusion SL2(𝐙)SL2(𝐑)\mathrm{SL}_{2}(\mathbf{Z})\leq\mathrm{SL}_{2}(\mathbf{R}) of the universal cover central extension

0𝐙SL~2(𝐑)SL2(𝐑)0.0\longrightarrow\mathbf{Z}\longrightarrow\widetilde{\mathrm{SL}}_{2}(\mathbf{R})\longrightarrow\mathrm{SL}_{2}(\mathbf{R})\longrightarrow 0.

Since the inclusion map SLd(𝐑)SLd+1(𝐑)\mathrm{SL}_{d}(\mathbf{R})\rightarrow\mathrm{SL}_{d+1}(\mathbf{R}) is surjective on fundamental groups for d2d\geq 2, the extension (7) agrees with the pushout of (8) along the quotient map 𝐙𝐙/2\mathbf{Z}\rightarrow\mathbf{Z}/2. Everything we need to know about these extensions, together with some useful information on the low-degree (co)homology of SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) is summarised in the following lemma.

Lemma 2.1.
  1. (i)

    The first two homology groups of SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) are given by the following table

    dd H1(SLd(𝐙);𝐙)\mathrm{H}_{1}(\mathrm{SL}_{d}(\mathbf{Z});\mathbf{Z}) H2(SLd(𝐙);𝐙)\mathrm{H}_{2}(\mathrm{SL}_{d}(\mathbf{Z});\mathbf{Z})
    22 𝐙/12\mathbf{Z}/12 0
    33 0 𝐙/2𝐙/2\mathbf{Z}/2\oplus\mathbf{Z}/2
    44 0 𝐙/2𝐙/2\mathbf{Z}/2\oplus\mathbf{Z}/2
    5\geq 5 0 𝐙/2\mathbf{Z}/2
  2. (ii)

    The map

    H2(SLd(𝐙);𝐙)H2(SLd+1(𝐙);𝐙)\mathrm{H}_{2}(\mathrm{SL}_{d}(\mathbf{Z});\mathbf{Z})\longrightarrow\mathrm{H}_{2}(\mathrm{SL}_{d+1}(\mathbf{Z});\mathbf{Z})

    induced by stabilisation is nontrivial for d3d\geq 3. For d=3d=3 its image has order two.

  3. (iii)

    The extension (8) is classified by a generator of

    H2(SL2(𝐙);𝐙)𝐙/12.\mathrm{H}^{2}(\mathrm{SL}_{2}(\mathbf{Z});\mathbf{Z})\cong\mathbf{Z}/12.
  4. (iv)

    The extension (7) is nontrivial for d2d\geq 2. For d5d\geq 5, it is classified by the generator of

    H2(SLd(𝐙);𝐙/2)𝐙/2.\mathrm{H}^{2}(\mathrm{SL}_{d}(\mathbf{Z});\mathbf{Z}/2)\cong\mathbf{Z}/2.
Proof.

That the abelianisation H1(SLd(𝐙);𝐙)\mathrm{H}_{1}(\mathrm{SL}_{d}(\mathbf{Z});\mathbf{Z}) is cyclic of order 1212 for d=2d=2 can be read off from the standard presentation of SL2(𝐙)\mathrm{SL}_{2}(\mathbf{Z}) (see e.g. [Mil71, Corollary 10.5]), and the fact that H1(SLd(𝐙);𝐙)\mathrm{H}_{1}(\mathrm{SL}_{d}(\mathbf{Z});\mathbf{Z}) vanishes for d3d\geq 3 follows for instance from [Mil71, Corollary 10.3] together with the observation that for d3d\geq 3 every elementary square matrix can be written as a commutator of two elementary matrices. The computation of H2(SLd(𝐙);𝐙)\mathrm{H}_{2}(\mathrm{SL}_{d}(\mathbf{Z});\mathbf{Z}) for d=2d=2 follows from the isomorphism SL2(𝐙)𝐙/4𝐙/2𝐙/6\mathrm{SL}_{2}(\mathbf{Z})\cong\mathbf{Z}/4\ast_{\mathbf{Z}/2}\mathbf{Z}/6 [Ser03, 1.5.3] and the Mayer–Vietoris sequence for group homology of amalgamated products [Bro94, Corollary II.7.7] (c.f. Exercise 3 on p. 52 of loc.cit.), for d=3,4d=3,4 from [vdK75], and for d5d\geq 5 from [Mil71, Corollary 5.8, Remark on p. 48, Theorem 10.1]. Using these calculations, the computations H2(SL2(𝐙);𝐙)𝐙/12\mathrm{H}^{2}(\mathrm{SL}_{2}(\mathbf{Z});\mathbf{Z})\cong\mathbf{Z}/12 and H2(SLd(𝐙);𝐙/2)𝐙/2\mathrm{H}^{2}(\mathrm{SL}_{d}(\mathbf{Z});\mathbf{Z}/2)\cong\mathbf{Z}/2 for d5d\geq 5 implicitly claimed in (iii) and (iv) follow from the universal coefficient theorem. The latter also implies the part of (ii) for d3d\geq 3 once we show (iv). The claim on the stabilising map for d=3d=3 can be proved via the arguments in [vdK75].

To prove (iii), we use the general fact that for a given central extension 0AEG00\rightarrow A\rightarrow E\rightarrow G\rightarrow 0, the Serre spectral sequence induces an exact sequence 0H1(G;A)H1(E;A)H1(A;A)H2(G;A)H2(E;A)0\rightarrow\mathrm{H}^{1}(G;A)\rightarrow\mathrm{H}^{1}(E;A)\rightarrow\mathrm{H}^{1}(A;A)\rightarrow\mathrm{H}^{2}(G;A)\rightarrow\mathrm{H}^{2}(E;A). The identity map induces a preferred class in H1(A;A)\mathrm{H}^{1}(A;A) and its image in H2(G;A)\mathrm{H}^{2}(G;A) is the class that classifies the given extension. Applying this to the extension (8), we see that in order to show that this extension generates H2(SL2(𝐙);𝐙)\mathrm{H}^{2}(\mathrm{SL}_{2}(\mathbf{Z});\mathbf{Z}) it suffices to show that H2(SL~2(𝐙);𝐙)\mathrm{H}^{2}(\smash{\widetilde{\mathrm{SL}}}_{2}(\mathbf{Z});\mathbf{Z}) vanishes. Now SL~2(𝐙)\smash{\widetilde{\mathrm{SL}}}_{2}(\mathbf{Z}) agrees up to isomorphism with the braid group B3B_{3} on three strands (see e.g. [Mil71, p. 83]), so the claim follows from the universal coefficient theorem and the facts that H1(B3;𝐙)𝐙\mathrm{H}_{1}(B_{3};\mathbf{Z})\cong\mathbf{Z} and H2(B3;𝐙)𝐙/2\mathrm{H}_{2}(B_{3};\mathbf{Z})\cong\mathbf{Z}/2 [Arn14, p. 32].

For (iv), we use the universal coefficient theorem to see that H2(SL2(𝐙);𝐙/2)𝐙/2\mathrm{H}^{2}(\mathrm{SL}_{2}(\mathbf{Z});\mathbf{Z}/2)\cong\mathbf{Z}/2 is surjected upon by the map H2(SLd(𝐙);𝐙)H2(SLd(𝐙);𝐙/2)\mathrm{H}^{2}(\mathrm{SL}_{d}(\mathbf{Z});\mathbf{Z})\rightarrow\mathrm{H}^{2}(\mathrm{SL}_{d}(\mathbf{Z});\mathbf{Z}/2) induced by reduction modulo 22, so it follows from (iii) that the extension (7) is nontrivial for d=2d=2 and hence also for all higher values of dd since the former is the pullback of the latter along the inclusion SL2(𝐙)SLd(𝐙)\mathrm{SL}_{2}(\mathbf{Z})\rightarrow\mathrm{SL}_{d}(\mathbf{Z}). As H2(SLd(𝐙);𝐙/2)𝐙/2\mathrm{H}^{2}(\mathrm{SL}_{d}(\mathbf{Z});\mathbf{Z}/2)\cong\mathbf{Z}/2 has only a single nontrivial element for d5d\geq 5, this gives (iv).∎

2.2. Mapping class groups of homotopy tori and D

We now determine the mapping class groups of homotopy tori of the form 𝒯=TdΣ\mathcal{T}=T^{d}\sharp\Sigma. The argument has three steps.

  1. Step \raisebox{-.9pt} {\arabic{subsubsection}}⃝

    Determine π0Diff+(TdΣ)\pi_{0}\,\mathrm{Diff}^{+}_{*}(T^{d}\sharp\Sigma) in terms of π0Diff((TdΣ))\pi_{0}\,\mathrm{Diff}_{\partial}((T^{d}\sharp\Sigma)^{\circ}).

  2. Step \raisebox{-.9pt} {\arabic{subsubsection}}⃝

    Determine π0Diff+(TdΣ)\pi_{0}\,\mathrm{Diff}^{+}_{*}(T^{d}\sharp\Sigma).

  3. Step \raisebox{-.9pt} {\arabic{subsubsection}}⃝

    Determine π0Diff+(TdΣ)\pi_{0}\,\mathrm{Diff}^{+}(T^{d}\sharp\Sigma).

Throughout this section, we abbreviate Td,:-(Td)T^{d,\circ}\coloneq(T^{d})^{\circ}, fix a basis of H1(Td,)𝐙d\mathrm{H}_{1}(T^{d,\circ})\cong\mathbf{Z}^{d}, and use the bases for the first homology groups of TdT^{d}, Td,Σ=(TdΣ)T^{d,\circ}\natural\Sigma^{\circ}=(T^{d}\sharp\Sigma)^{\circ}, and TdΣT^{d}\sharp\Sigma that are induced by the chosen basis of H1(Td,)\mathrm{H}_{1}(T^{d,\circ}).

Step \raisebox{-.9pt} {\arabic{subsubsection}}⃝. Fixing a disc or a point

We first determine the group π0Diff(𝒯)\pi_{0}\,\mathrm{Diff}_{\partial}(\mathcal{T}^{\circ}) in terms of the group π0Diff+(𝒯)\pi_{0}\,\mathrm{Diff}^{+}_{\ast}(\mathcal{T}). This step works for general homotopy tori 𝒯\mathcal{T}, not just those of the form TdΣT^{d}\sharp\Sigma.

Lemma 2.2.

For a homotopy dd-torus 𝒯\mathcal{T}, there are pullback squares

π0Diff(𝒯){\pi_{0}\,\mathrm{Diff}_{\partial}(\mathcal{T}^{\circ})}π0Diff+(𝒯){\pi_{0}\,\mathrm{Diff}^{+}_{\ast}(\mathcal{T})}SL~d(𝐙){\widetilde{\mathrm{SL}}_{d}(\mathbf{Z})}SLd(𝐙),{\mathrm{SL}_{d}(\mathbf{Z}),}ext\scriptstyle{\mathrm{ext}}\scriptstyle{\cong}\scriptstyle{\cong}for d=2d=2 andπ0Diff(𝒯){\pi_{0}\,\mathrm{Diff}_{\partial}(\mathcal{T}^{\circ})}π0Diff+(𝒯){\pi_{0}\,\mathrm{Diff}^{+}_{\ast}(\mathcal{T})}SL¯d(𝐙){\overline{\mathrm{SL}}_{d}(\mathbf{Z})}SLd(𝐙){\mathrm{SL}_{d}(\mathbf{Z})}ext\scriptstyle{\mathrm{ext}}   for d3d\geq 3
Proof.

If d=2d=2, then 𝒯\mathcal{T} is the standard 22-torus T2T^{2} for which the claimed square is well-known (for a reference, compare the standard presentations of π0Diff(Td,)\pi_{0}\,\mathrm{Diff}_{\partial}(T^{d,\circ}) and SL~d(𝐙)\widetilde{\mathrm{SL}}_{d}(\mathbf{Z}) e.g. in [Mil71, p.82–83] and [Kor02, Section 5]). For d3d\geq 3, we consider the map of central extensions

(9) 0{0}π1SO(d){\pi_{1}\,\mathrm{SO}(d)}π0Diff(𝒯){\pi_{0}\,\mathrm{Diff}_{\partial}(\mathcal{T}^{\circ})}π0Diff+(𝒯){\pi_{0}\,\mathrm{Diff}^{+}_{\ast}(\mathcal{T})}0{0}0{0}π1STop{\pi_{1}\,\mathrm{STop}}π0Homeo~TTd(Td){\pi_{0}\,\smash{\mathrm{\widetilde{Homeo}}}_{T_{\ast}T^{d}}(T^{d})}π0Homeo~+(Td){\pi_{0}\,\smash{\mathrm{\widetilde{Homeo}}}^{+}_{\ast}(T^{d})}0{0}\scriptstyle{\cong}

induced by (6) for M=𝒯M=\mathcal{T}, using that the bottom sequence only depends on the underlying topological manifold. Exactness at π1STop\pi_{1}\,\mathrm{STop} follows from the fact that π1Homeo~+(Td)=0\pi_{1}\,\smash{\mathrm{\widetilde{Homeo}}}^{+}_{\ast}(T^{d})=0 by 1.11 and exactness at π1SO(d)\pi_{1}\,\mathrm{SO}(d) follows from exactness at π1STop\pi_{1}\,\mathrm{STop}. 1.11 also shows that the homology action map π0Homeo~+(Td)SLd(𝐙)\pi_{0}\,\smash{\mathrm{\widetilde{Homeo}}}^{+}_{\ast}(T^{d})\rightarrow\mathrm{SL}_{d}(\mathbf{Z}) is an isomorphism, so we are left to show that the bottom extension is isomorphic to 0𝐙/2SL¯d(𝐙)SLd(𝐙)0\smash{0\rightarrow\mathbf{Z}/2\rightarrow\overline{\mathrm{SL}}_{d}(\mathbf{Z})\rightarrow\mathrm{SL}_{d}(\mathbf{Z})\rightarrow 0}. It suffices to show this for large enough dd, since the bottom extension in dimension dd maps by taking products with S1S^{1} to the corresponding extension in dimension d+1d+1 (which have both kernel π1STop\pi_{1}\mathrm{STop}), so the extension for dd is the pullback of the extension for d+1d+1 along the inclusion SLd(𝐙)SLd+1(𝐙)\mathrm{SL}_{d}(\mathbf{Z})\rightarrow\mathrm{SL}_{d+1}(\mathbf{Z}). We may thus assume d5d\geq 5 in which case there is a single nontrivial central extension of SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) by 𝐙/2\mathbf{Z}/2 (see 2.1), so we only need to exclude that the bottom extension in (9) is trivial. To show this, we consider (9) for 𝒯=Td\mathcal{T}=T^{d} and extend it to the top as

0{0}π1SO(2){\pi_{1}\,\mathrm{SO}(2)}π0Diff(T2,){\pi_{0}\,\mathrm{Diff}_{\partial}(T^{2,\circ})}π0Diff+(T2){\pi_{0}\,\mathrm{Diff}^{+}_{\ast}(T^{2})}0{0}0{0}π1SO(d){\pi_{1}\,\mathrm{SO}(d)}π0Diff(Td,){\pi_{0}\,\mathrm{Diff}_{\partial}(T^{d,\circ})}π0Diff+(Td){\pi_{0}\,\mathrm{Diff}^{+}_{\ast}(T^{d})}0{0}0{0}π1STop{\pi_{1}\,\mathrm{STop}}π0Homeo~TTd(Td){\pi_{0}\,\smash{\mathrm{\widetilde{Homeo}}}_{T_{\ast}T^{d}}(T^{d})}π0Homeo~+(Td){\pi_{0}\,\smash{\mathrm{\widetilde{Homeo}}}^{+}_{\ast}(T^{d})}0.{0.}inc\scriptstyle{\mathrm{inc}_{*}}\scriptstyle{\cong}

where the top middle vertical map is induced by taking products with idTd2\mathrm{id}_{T^{d-2}} followed by restriction, commutativity of the left upper square follows by an application of 1.1 (i) to M=T2M=T^{2} and N=Td2N=T^{d-2}, and the right upper square is induced by the commutativity of the left upper square. We will show that the bottom extension is nontrivial by showing that its pullback along the composition π0Diff+(T2)π0Homeo~+(Td)\pi_{0}\,\mathrm{Diff}^{+}_{\ast}(T^{2})\rightarrow\pi_{0}\,\smash{\mathrm{\widetilde{Homeo}}}^{+}_{\ast}(T^{d}) is nontrivial. This composition is isomorphic to the inclusion SL2(𝐙)SLd(𝐙)\mathrm{SL}_{2}(\mathbf{Z})\rightarrow\mathrm{SL}_{d}(\mathbf{Z}) and the composition π1SO(2)π1STop\pi_{1}\,\mathrm{SO}(2)\rightarrow\pi_{1}\,\mathrm{STop} to the quotient map 𝐙𝐙/2\mathbf{Z}\rightarrow\mathbf{Z}/2, so it follows that the pullback in question is isomorphic to the mod 22 reduction of the extension 0𝐙SL~2(𝐙)SL2(𝐙)0\smash{0\rightarrow\mathbf{Z}\rightarrow\widetilde{\mathrm{SL}}_{2}(\mathbf{Z})\rightarrow\mathrm{SL}_{2}(\mathbf{Z})\rightarrow 0}, i.e. the extension 0𝐙/2SL¯2(𝐙)SL2(𝐙)0\smash{0\rightarrow\mathbf{Z}/2\rightarrow\overline{\mathrm{SL}}_{2}(\mathbf{Z})\rightarrow\mathrm{SL}_{2}(\mathbf{Z})\rightarrow 0}. The latter is nontrivial by 2.1 (iv).∎

Step \raisebox{-.9pt} {\arabic{subsubsection}}⃝. The pointed mapping class group of TdΣT^{d}\sharp\Sigma

Next, we determine the group π0Diff+(TdΣ)\pi_{0}\,\mathrm{Diff}^{+}_{*}(T^{d}\sharp\Sigma). For Σ=Sd\Sigma=S^{d} the evaluation fibration Diff+(Td)Td\mathrm{Diff}^{+}(T^{d})\rightarrow T^{d} whose fibre is Diff+(Td)\mathrm{Diff}^{+}_{*}(T^{d}) has a splitting given by the standard action of TdT^{d} on itself, so the long exact sequence in homotopy groups induces the first out of two isomorphisms

π0Diff+(Td)π0Diff+(Td)d6SLd(𝐙)Ω;\pi_{0}\,\mathrm{Diff}^{+}_{*}(T^{d})\cong\pi_{0}\,\mathrm{Diff}^{+}(T^{d})\overset{d\geq 6}{\cong}\mathrm{SL}_{d}(\mathbf{Z})\ltimes\Omega;

the second isomorphism was explained in the introduction. Combining this with 2.2 for 𝒯=Td\mathcal{T}=T^{d}, we obtain an isomorphism

(10) π0Diff(Td,)SL¯d(𝐙)Ωfor d6.\pi_{0}\,\mathrm{Diff}_{\partial}(T^{d,\circ})\cong\overline{\mathrm{SL}}_{d}(\mathbf{Z})\ltimes\Omega\quad\text{for }d\geq 6.

Now recall that the collar twist tTdπ0Diff(Td,)t_{T^{d}}\in\pi_{0}\,\mathrm{Diff}_{\partial}(T^{d,\circ}) generates the kernel of the map to π0Diff+(Td)\pi_{0}\,\mathrm{Diff}^{+}_{\ast}(T^{d}) so it corresponds under the isomorphism (10) to the element (td,0)(t_{d},0) where tdSL¯d(𝐙)t_{d}\in\overline{\mathrm{SL}}_{d}(\mathbf{Z}) is the central element that generates the kernel of the map to SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}). The composition

(11) Θd+1π0Diff(Dd)idTd,()π0Diff(Td,),\Theta_{d+1}\cong\pi_{0}\,\mathrm{Diff}_{\partial}(D^{d})\xrightarrow{\mathrm{id}_{T^{d,\circ}}\natural(-)}\pi_{0}\,\mathrm{Diff}_{\partial}(T^{d,\circ}),

which we abbreviate by ιd:Θd+1π0Diff(Td,)\iota_{d}\colon\Theta_{d+1}\rightarrow\pi_{0}\,\mathrm{Diff}_{\partial}(T^{d,\circ}), can be identified in terms of (10) as follows:

Lemma 2.3.

With respect to the isomorphism (10), the composition (11) agrees with the inclusion Θd+1SL¯d(𝐙)Ω\Theta_{d+1}\leq\overline{\mathrm{SL}}_{d}(\mathbf{Z})\ltimes\Omega given by the (j=0)(j=0)-summand in (1).

Proof.

By an application of 2.2, this would follow from showing the analogous statement for π0Diff+(Td)SLd(𝐙)Ω\pi_{0}\,\mathrm{Diff}^{+}_{*}(T^{d})\cong\mathrm{SL}_{d}(\mathbf{Z})\ltimes\Omega instead of for π0Diff(Td,)\pi_{0}\,\mathrm{Diff}_{\partial}(T^{d,\circ}) once we know that the postcomposition of (11) with the isomorphism (10) has image in ΩSL¯d(𝐙)Ω\Omega\leq\overline{\mathrm{SL}}_{d}(\mathbf{Z})\ltimes\Omega. That the statement holds in π0Diff+(Td)π0Diff+(Td)\pi_{0}\,\mathrm{Diff}^{+}_{*}(T^{d})\cong\pi_{0}\,\mathrm{Diff}^{+}(T^{d}) follows from [Hat83, p. 9, Remark (5)], so it suffices to show that (11) lands in the subgroup Ω\Omega. To show that, note that diffeomorphisms in the image of (11) are topologically isotopic to the identity since Homeo(Dd)\mathrm{Homeo}_{\partial}(D^{d}) is contractible by the Alexander trick. In particular, (11) lands in the kernel of the forgetful map π0Diff(Td,)π0Homeo~TTd(Td)\pi_{0}\mathrm{Diff}_{\partial}(T^{d,\circ})\rightarrow\pi_{0}\,\smash{\mathrm{\widetilde{Homeo}}}_{T_{\ast}T^{d}}(T^{d}) which agrees via the isomorphism π0Diff+(Td)π0Diff+(Td)SLd(𝐙)Ω\pi_{0}\,\mathrm{Diff}^{+}_{*}(T^{d})\cong\pi_{0}\,\mathrm{Diff}^{+}(T^{d})\cong\mathrm{SL}_{d}(\mathbf{Z})\ltimes\Omega precisely with the subgroup Ω\Omega (see the proof of 2.2), so the claim follows. ∎

Given a homotopy sphere ΣΘd\Sigma\in\Theta_{d} and d7d\geq 7, we have the isomorphism discussed in Section 1.3

(12) π0Diff(Td,)()idΣπ0Diff((TdΣ)),\pi_{0}\,\mathrm{Diff}_{\partial}(T^{d,\circ})\overset{(-)\natural\mathrm{id}_{\Sigma^{\circ}}}{\cong}\pi_{0}\,\mathrm{Diff}_{\partial}((T^{d}\sharp\Sigma)^{\circ}),

so from the exact sequence (3), we see that π0Diff(TdΣ)\pi_{0}\,\mathrm{Diff}_{*}(T^{d}\sharp\Sigma) is isomorphic to the quotient of π0Diff(Td,)SL¯d(𝐙)Ω\smash{\pi_{0}\,\mathrm{Diff}_{\partial}(T^{d,\circ})\cong\overline{\mathrm{SL}}_{d}(\mathbf{Z})\ltimes\Omega} by the central subgroup generated by the preimage of the collar twist tTdΣπ0Diff((TdΣ))t_{T^{d}\sharp\Sigma}\in\pi_{0}\,\mathrm{Diff}_{\partial}((T^{d}\sharp\Sigma)^{\circ}) under (12).

Lemma 2.4.

The preimage of tTdΣπ0Diff((TdΣ))\smash{t_{T^{d}\sharp\Sigma}\in\pi_{0}\mathrm{Diff}_{\partial}((T^{d}\sharp\Sigma)^{\circ})} in SL¯d(𝐙)Ω\overline{\mathrm{SL}}_{d}(\mathbf{Z})\ltimes\Omega under the combined isomorphisms (12) and (10) is (td,ηΣ)SL¯d(𝐙)Ω(t_{d},\eta\cdot\Sigma)\in\overline{\mathrm{SL}}_{d}(\mathbf{Z})\ltimes\Omega. Consequently, we have an isomorphism

(13) π0Diff+(TdΣ)(SL¯d(𝐙)Ω)/(td,ηΣ).\pi_{0}\,\mathrm{Diff}^{+}_{\ast}(T^{d}\sharp\Sigma)\cong\big{(}\overline{\mathrm{SL}}_{d}(\mathbf{Z})\ltimes\Omega\big{)}/\langle(t_{d},\eta\cdot\Sigma)\rangle.

for d7d\geq 7 which is compatible with the homomorphisms to SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}).

Proof.

We already explained how the second part follows from the first. To prove the first, we use the relation tTdΣ=tTdidΣ+idTd,tΣ\smash{t_{T^{d}\sharp\Sigma}=t_{T^{d}}\natural\mathrm{id}_{\Sigma^{\circ}}+\mathrm{id}_{T^{d,\circ}}\natural t_{\Sigma}} in π0Diff((TdΣ))\pi_{0}\,\mathrm{Diff}_{\partial}((T^{d}\sharp\Sigma)^{\circ}) ensured by 1.1 (ii), using which we express the element in question in π0Diff(Td,)\pi_{0}\,\mathrm{Diff}_{\partial}(T^{d,\circ}) as

(tTdidΣ+idTd,tΣ)idΣ¯=tTd+idTd,(tΣidΣ¯)=tTd+ιd(ηΣ)π0Diff(Td,).\big{(}t_{T^{d}}\natural\mathrm{id}_{\Sigma^{\circ}}+\mathrm{id}_{T^{d,\circ}}\natural t_{\Sigma}\big{)}\natural\mathrm{id}_{\overline{\Sigma}^{\circ}}=t_{T^{d}}+\mathrm{id}_{T^{d,\circ}}\natural(t_{\Sigma}\natural\mathrm{id}_{\overline{\Sigma}^{\circ}})=t_{T^{d}}+\iota_{d}(\eta\cdot\Sigma)\in\pi_{0}\,\mathrm{Diff}_{\partial}(T^{d,\circ}).

Here we used the equality TΣ=ηΣT_{\Sigma}=\eta\cdot\Sigma from 1.5 and the definition of TΣT_{\Sigma} from Section 1.3. By the discussion above, tTdt_{T^{d}} and ιd(ηΣ)\iota_{d}(\eta\cdot\Sigma) corresponds under the isomorphism (10) to the elements (td,0)(t_{d},0) and (0,ηΣ)(0,\eta\cdot\Sigma) in SL¯d(𝐙)Ω\smash{\overline{\mathrm{SL}}_{d}(\mathbf{Z})\ltimes\Omega}, so the element we are looking for is indeed (td,ηΣ)(t_{d},\eta\cdot\Sigma).∎

The quotient of SL¯d(𝐙)Ω\overline{\mathrm{SL}}_{d}(\mathbf{Z})\ltimes\Omega appearing in 2.4 can be further simplified:

Lemma 2.5.

There is an isomorphism of groups

(SL¯d(𝐙)Ω)/(td,ηΣ){SL¯d(𝐙)(Ω/ηΣ) if ηΣΘd+1is not divisible by 2SLd(𝐙)Ω if ηΣΘd+1is divisible by 2\big{(}\overline{\mathrm{SL}}_{d}(\mathbf{Z})\ltimes\Omega\big{)}/\langle(t_{d},\eta\cdot\Sigma)\rangle\cong\begin{cases}\overline{\mathrm{SL}}_{d}(\mathbf{Z})\ltimes\big{(}\Omega/\langle\eta\cdot\Sigma\rangle\big{)}&\text{ if }\eta\cdot\Sigma\in\Theta_{d+1}\text{is not divisible by }2\\ \mathrm{SL}_{d}(\mathbf{Z})\ltimes\Omega&\text{ if }\eta\cdot\Sigma\in\Theta_{d+1}\text{is divisible by }2\\ \end{cases}

that is compatible with the homomorphisms to SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}).

Proof.

Since the element ηΣ\eta\cdot\Sigma of the finite abelian group Θd+1\Theta_{d+1} is of order 22, it is not divisible by 22 if and only if it generates a direct 𝐙/2\mathbf{Z}/2-summand. We first assume that this is the case, so Θd+1(Θd+1/ηΣ)𝐙/2\Theta_{d+1}\cong(\Theta_{d+1}/\langle\eta\cdot\Sigma\rangle)\oplus\mathbf{Z}/2. Writing ΩΩ\Omega^{\prime}\leq\Omega for the SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z})-invariant subgroup complementary to the central summand Θd+1Ω\Theta_{d+1}\leq\Omega in (1), we have SL¯d(𝐙)ΩSL¯d(𝐙)(Ω(Θd/ηΣ)𝐙/2)\smash{\overline{\mathrm{SL}}_{d}(\mathbf{Z})\ltimes\Omega}\cong\smash{\overline{\mathrm{SL}}_{d}(\mathbf{Z})\ltimes(\Omega^{\prime}\oplus(\Theta_{d}/\langle\eta\cdot\Sigma\rangle)\oplus\mathbf{Z}/2)}. The latter admits an epimorphism to SL¯d(𝐙)(Ω(Θd/ηΣ))\smash{\overline{\mathrm{SL}}_{d}(\mathbf{Z})\ltimes(\Omega^{\prime}\oplus(\Theta_{d}/\langle\eta\cdot\Sigma\rangle))} given by sending (A,a+Σ+x)(A,a+\Sigma^{\prime}+x) to (xtdA,a+Σ)(x\cdot t_{d}\cdot A,a+\Sigma^{\prime}). This is well-defined since the central element tdSL¯d(𝐙)\smash{t_{d}\in\overline{\mathrm{SL}}_{d}(\mathbf{Z})} has order 22 and acts trivially on Ω\Omega since the action factors by construction through SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}). The kernel of this epimorphism is the subgroup generated by (td,ηΣ)(t_{d},\eta\cdot\Sigma), so we obtain an isomorphism between (SL¯d(𝐙)Ω)/(td,ηΣ)\smash{\big{(}\overline{\mathrm{SL}}_{d}(\mathbf{Z})\ltimes\Omega\big{)}/\langle(t_{d},\eta\cdot\Sigma)\rangle} and SL¯d(𝐙)(Ω/ηΣ))\overline{\mathrm{SL}}_{d}(\mathbf{Z})\ltimes\big{(}\Omega/\langle\eta\cdot\Sigma\rangle)\big{)}, as claimed.

Now assume that ηΣΘd+1\eta\cdot\Sigma\in\Theta_{d+1} does not generate a direct summand, so is divisible by 22. We have a (non-central) extension

(14) 0(Ω𝐙/2)/ηΣ+[1](SL¯d(𝐙)Ω)/(td,ηΣ)SLd(𝐙)0.0\longrightarrow\big{(}\Omega\oplus\mathbf{Z}/2\big{)}/\langle\eta\cdot\Sigma+[1]\rangle\longrightarrow\big{(}\overline{\mathrm{SL}}_{d}(\mathbf{Z})\ltimes\Omega\big{)}/\langle(t_{d},\eta\cdot\Sigma)\rangle\longrightarrow\mathrm{SL}_{d}(\mathbf{Z})\longrightarrow 0.

As ηΣ\eta\cdot\Sigma has order 22, the SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z})-equivariant map from Ω=Ω0\Omega=\Omega\oplus 0 into the kernel of (14) induced by inclusion is an isomorphism, so in order to show that (SL¯d(𝐙)Ω)/(td,ηΣ)\smash{\big{(}\overline{\mathrm{SL}}_{d}(\mathbf{Z})\ltimes\Omega\big{)}/\langle(t_{d},\eta\cdot\Sigma)\rangle} is isomorphic to SLd(𝐙)Ω\mathrm{SL}_{d}(\mathbf{Z})\ltimes\Omega it suffices to show that (14) splits. Writing Ω=Θd+1Ω\Omega=\Theta_{d+1}\oplus\Omega^{\prime} as in the previous case, the extension (14) is by construction the sum of the trivial SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z})-extension by the SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z})-module Ω\Omega^{\prime} with the central extension classified by the image of the unique nontrivial element in H2(SLd(𝐙);𝐙/2)\mathrm{H}^{2}(\mathrm{SL}_{d}(\mathbf{Z});\mathbf{Z}/2) under the composition H2(SLd(𝐙);𝐙/2)H2(SLd(𝐙);𝐙/2Θd+1)H2(SLd(𝐙);(𝐙/2Θd+1)/(td,ηΣ))\mathrm{H}^{2}(\mathrm{SL}_{d}(\mathbf{Z});\mathbf{Z}/2)\rightarrow\mathrm{H}^{2}(\mathrm{SL}_{d}(\mathbf{Z});\mathbf{Z}/2\oplus\Theta_{d+1})\rightarrow\mathrm{H}^{2}(\mathrm{SL}_{d}(\mathbf{Z});(\mathbf{Z}/2\oplus\Theta_{d+1})/(t_{d},\eta\cdot\Sigma)) induced by the inclusion and quotient maps of coefficients, so it suffices to show that this image is trivial. From the universal coefficient theorem and the computations in 2.1, we see that for any abelian group, the map H2(SLd(𝐙);A)H2(SLd(𝐙);A/2)\mathrm{H}^{2}(\mathrm{SL}_{d}(\mathbf{Z});A)\rightarrow\mathrm{H}^{2}(\mathrm{SL}_{d}(\mathbf{Z});A/2) induced by reducing modulo 22 is an isomorphism, so to show that the class in question is trivial, it suffices to do so after reducing modulo 22. The latter follows by noting that the composition of SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z})-modules 𝐙/2𝐙/2Θd+1(𝐙/2Θd+1)/(td,ηΣ))\mathbf{Z}/2\subset\mathbf{Z}/2\oplus\Theta_{d+1}\rightarrow\big{(}\mathbf{Z}/2\oplus\Theta_{d+1})/(t_{d},\eta\cdot\Sigma)\big{)} is trivial after passing to (𝐙/2Θd+1)/(td,ηΣ))/2\big{(}\mathbf{Z}/2\oplus\Theta_{d+1})/(t_{d},\eta\cdot\Sigma)\big{)}/2 since ηΣ\eta\cdot\Sigma vanishes in Θd+1/2\Theta_{d+1}/2 by assumption. ∎

Step \raisebox{-.9pt} {\arabic{subsubsection}}⃝. Fixing a point or not

Using the description of π0Diff+(TdΣ)\pi_{0}\,\mathrm{Diff}^{+}_{*}(T^{d}\sharp\Sigma) from the previous step, we are now in the position to determine π0Diff+(TdΣ)\pi_{0}\,\mathrm{Diff}^{+}(T^{d}\sharp\Sigma). In view of the fibration sequence

(15) TdΣBDiff+(TdΣ)BDiff+(TdΣ)T^{d}\sharp\Sigma\longrightarrow\mathrm{BDiff}^{+}_{*}(T^{d}\sharp\Sigma)\longrightarrow\mathrm{BDiff}^{+}(T^{d}\sharp\Sigma)

this amounts to understanding the image of the “point-pushing” homomorphism p:π1TdΣπ0Diff+(TdΣ)\mathrm{p}\colon\pi_{1}\,T^{d}\sharp\Sigma\to\pi_{0}\,\mathrm{Diff}^{+}_{*}(T^{d}\sharp\Sigma). For Σ=Sd\Sigma=S^{d}, the image is trivial a result of the action of TdT^{d} on itself (see the beginning of Step \raisebox{-.9pt} {\arabic{subsubsection}}⃝), but for any other homotopy sphere such an action is not available and it in fact follows from the following lemma that the image is never trivial.

Lemma 2.6.

For d7d\geq 7 and ΣΘd\Sigma\in\Theta_{d}, the map

𝐙d=π1TdΣpπ0Diff+(TdΣ)\mathbf{Z}^{d}=\pi_{1}\,T^{d}\sharp\Sigma\overset{\mathrm{p}}{\longrightarrow}\pi_{0}\,\mathrm{Diff}^{+}_{*}(T^{d}\sharp\Sigma)

agrees with the composition

𝐙d()Σ𝐙dΘdSL¯d(𝐙)Ωπ0Diff(Td,)()idΣπ0Diff((TdΣ))extπ0Diff+(TdΣ)\mathbf{Z}^{d}\overset{(-)\otimes\Sigma}{\longrightarrow}\mathbf{Z}^{d}\otimes\Theta_{d}\leq\overline{\mathrm{SL}}_{d}(\mathbf{Z})\ltimes\Omega\cong\pi_{0}\,\mathrm{Diff}_{\partial}(T^{d,\circ})\overset{(-)\natural\mathrm{id}_{\Sigma^{\circ}}}{\cong}\pi_{0}\,\mathrm{Diff}_{\partial}((T^{d}\sharp\Sigma)^{\circ})\overset{\mathrm{ext}}{\longrightarrow}\pi_{0}\,\mathrm{Diff}^{+}_{*}(T^{d}\sharp\Sigma)

involving the isomorphism π0Diff((TdΣ))SL¯d(𝐙)Ω\pi_{0}\,\mathrm{Diff}_{\partial}((T^{d}\sharp\Sigma)^{\circ})\cong\overline{\mathrm{SL}}_{d}(\mathbf{Z})\ltimes\Omega from (10).

Proof.

It suffices to show that the two compositions agree on the first standard basis vector e1𝐙de_{1}\in\mathbf{Z}^{d}, since both compositions are π0Diff(Td,)\pi_{0}\,\mathrm{Diff}_{\partial}(T^{d,\circ})-equivariant (the action on the source is through SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) and the action on the target is by conjugation after extending diffeomorphisms from Td,T^{d,\circ} to TdΣT^{d}\sharp\Sigma by the identity) and the orbit of e1𝐙de_{1}\in\mathbf{Z}^{d} under the SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z})-action spans 𝐙d\mathbf{Z}^{d}.

We work with the following model of TdΣT^{d}\sharp\Sigma: view Td1T^{d-1} as 𝐑d1/𝐙d1\mathbf{R}^{d-1}/\mathbf{Z}^{d-1}, choose an orientation-preserving embedding ι:Dd1Td1=𝐑d1/𝐙d1\iota\colon D^{d-1}\hookrightarrow T^{d-1}=\mathbf{R}^{d-1}/\mathbf{Z}^{d-1} disjoint from the origin [0]𝐑d1/𝐙d1=Td1[0]\in\mathbf{R}^{d-1}/\mathbf{Z}^{d-1}=T^{d-1} and a representative fΣDiff(Dd1)f_{\Sigma}\in\mathrm{Diff}_{\partial}(D^{d-1}) of Σπ0Diff(Dd1)Θd\Sigma\in\pi_{0}\,\mathrm{Diff}_{\partial}(D^{d-1})\cong\Theta_{d}, extend fΣf_{\Sigma} by the identity to a diffeomorphism FΣDiff+(Td1)F_{\Sigma}\in\mathrm{Diff}^{+}(T^{d-1}) supported in int(ι(Dd1))\mathrm{int}(\iota(D^{d-1})), and form the mapping torus TdΣ:-([0,1]×Td1)/((1,x)(0,FΣ(x))T^{d}\sharp\Sigma\coloneq([0,1]\times T^{d-1})/((1,x)\sim(0,F_{\Sigma}(x)). We parametrise this quotient by [0,1)×Td1[0,1)\times T^{d-1} in the evident way, use [0,0]TdΣ[0,0]\in T^{d}\sharp\Sigma as base point, and we view Td,T^{d,\circ} as the complement of an embedded disc DdTdΣD^{d}\subset T^{d}\sharp\Sigma that contains the part where the nontrivial gluing happened (i.e. the image of {0}×ι(Dd1)\{0\}\times\iota(D^{d-1}) in the quotient) and is disjoint from the image of [0,1]×{0}[0,1]\times\{0\} in the quotient. The latter is so that the loop ([(t,0)])t[0,1]([(t,0)])_{t\in[0,1]} in TdΣT^{d}\sharp\Sigma is contained in Td,T^{d,\circ}. We chose a basis for H1(Td,)\mathrm{H}_{1}(T^{d,\circ}) such that this loop represents e1𝐙de_{1}\in\mathbf{Z}^{d}.

The first claim we show is that the image of e1𝐙de_{1}\in\mathbf{Z}^{d} under the map p:π1TdΣπ0Diff+(TdΣ)\mathrm{p}\colon\pi_{1}\,T^{d}\sharp\Sigma\rightarrow\pi_{0}\,\mathrm{Diff}^{+}_{*}(T^{d}\sharp\Sigma) is represented by the diffeomorphism ϕDiff+(TdΣ)\phi\in\mathrm{Diff}^{+}_{*}(T^{d}\sharp\Sigma) given by using id×fΣ\mathrm{id}\times f_{\Sigma} on the image of id×ι\mathrm{id}\times\iota in TdΣT^{d}\sharp\Sigma, and extending it to all of TdΣT^{d}\sharp\Sigma by the identity. This is because being the connecting map in the long exact sequence induced by the evaluation fibration ev[0,0]:Diff+(TdΣ)TdΣ\mathrm{ev}_{[0,0]}\colon\mathrm{Diff}^{+}(T^{d}\sharp\Sigma)\rightarrow T^{d}\sharp\Sigma with fibre Diff+(TdΣ)\mathrm{Diff}_{\ast}^{+}(T^{d}\sharp\Sigma), the point-pushing map sends e1𝐙de_{1}\in\mathbf{Z}^{d} to the isotopy class of any diffeomorphism ϕ1\phi_{1} that arises as the value at time t=1t=1 of a path (ϕt)t[0,1](\phi_{t})_{t\in[0,1]} in Diff+(TdΣ)\mathrm{Diff}^{+}(T^{d}\sharp\Sigma) with ϕ0=id\phi_{0}=\mathrm{id} and ϕt([0,0])=[t,0]\phi_{t}([0,0])=[t,0]. A possible choice of such path is given by ϕt([s,x]):-[s+t,x]\phi_{t}([s,x])\coloneq[s+t,x] for s+t<1s+t<1 and ϕt([s,x]):-[s+t1,FΣ(x)]\phi_{t}([s,x])\coloneq[s+t-1,F_{\Sigma}(x)] for s+t1s+t\geq 1, which indeed agrees with ϕ\phi at time 11.

The second claim we make is that the image of e1𝐙de_{1}\in\mathbf{Z}^{d} under the second composition in the statement is given by the diffeomorphism obtained by choosing an orientation-preserving embedding ι:Dd1Td1\iota^{\prime}\colon D^{d-1}\hookrightarrow T^{d-1} such that the image of id×ι\mathrm{id}\times\iota^{\prime} in TdΣT^{d}\sharp\Sigma is contained in Td,T^{d,\circ} and avoids the origin, using id×fΣ\mathrm{id}\times f_{\Sigma} on the image of id×ι\mathrm{id}\times\iota^{\prime} in TdΣT^{d}\sharp\Sigma and extending it to a diffeomorphism of TdΣT^{d}\sharp\Sigma by the identity. This would imply the result, since the image of e1e_{1} under both maps in consideration arises from the following construction: choose an embedding S1×Dd1TdΣ\{[0,0]}S^{1}\times D^{d-1}\hookrightarrow T^{d}\sharp\Sigma\backslash\{[0,0]\} that represents e1𝐙d=H1(TdΣ)e_{1}\in\mathbf{Z}^{d}=\mathrm{H}_{1}(T^{d}\sharp\Sigma) (which is unique up to isotopy as d4d\geq 4), use id×fΣ\mathrm{id}\times f_{\Sigma} on this image, and extend by the identity.

To show this claim, we prove more generally that the composition 𝐙dΘdΩSL¯d(𝐙)Ωπ0Diff(Td,)\mathbf{Z}^{d}\otimes\Theta_{d}\leq\Omega\leq\overline{\mathrm{SL}}_{d}(\mathbf{Z})\ltimes\Omega\cong\pi_{0}\,\mathrm{Diff}_{\partial}(T^{d,\circ}) is given by sending xΣ𝐙dΘdx\otimes\Sigma^{\prime}\in\mathbf{Z}^{d}\otimes\Theta_{d} to the diffeomorphism obtained by representing xπ1Td,x\in\pi_{1}\,T^{d,\circ} by an embedding S1×Dd1Td,S^{1}\times D^{d-1}\subset T^{d,\circ} and ΣΘdπ0Diff(Dd)\Sigma^{\prime}\in\Theta_{d}\cong\pi_{0}\,\mathrm{Diff}_{\partial}(D^{d}) by a diffeomorphism fΣDiff(Dd)f_{\Sigma^{\prime}}\in\mathrm{Diff}_{\partial}(D^{d}), using id×fΣ\mathrm{id}\times f_{\Sigma^{\prime}} on S1×Dd1S^{1}\times D^{d-1} and extending it to TdT^{d} by the identity. By the argument from the proof of 2.3, it suffices to show that the described diffeomorphism considered as a diffeomorphism of π0Diff(Td)SLd(𝐙)Ω\pi_{0}\,\mathrm{Diff}_{*}(T^{d})\cong\mathrm{SL}_{d}(\mathbf{Z})\ltimes\Omega agrees with the image of e1Σe_{1}\otimes\Sigma^{\prime} under the inclusion 𝐙dΘdΩ\mathbf{Z}^{d}\otimes\Theta_{d}\leq\Omega. This follows from [Hat78, p. 9, Remark (5)]. ∎

Combining 2.6 with 2.4, the long exact sequence induced by (15) implies:

Corollary 2.7.

For d7d\geq 7 and ΣΘd\Sigma\in\Theta_{d} there is an isomorphism

π0Diff+(TdΣ)(SL¯d(𝐙)Ω)/((td,ηΣ)(𝐙dΣ))\pi_{0}\,\mathrm{Diff}^{+}(T^{d}\sharp\Sigma)\cong\big{(}\overline{\mathrm{SL}}_{d}(\mathbf{Z})\ltimes\Omega\big{)}/\big{(}\langle(t_{d},\eta\cdot\Sigma)\rangle\oplus(\mathbf{Z}^{d}\otimes\langle\Sigma\rangle)\big{)}

which is compatible with the homomorphisms to SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}).

From this, the asserted identification of π0Diff+(TdΣ)\pi_{0}\,\mathrm{Diff}^{+}(T^{d}\sharp\Sigma) in D follows by proving that the right-hand quotient in the previous corollary can be simplified to the semidirect product

{SL¯d(𝐙)[Ω/(ηΣ(𝐙dΣ)]if ηΣΘd+1is not divisible by 2SLd(𝐙)[Ω/(𝐙dΣ)]if ηΣΘd+1 is divisible by 2\begin{cases}\overline{\mathrm{SL}}_{d}(\mathbf{Z})\ltimes\Big{[}\Omega/\big{(}\langle\eta\cdot\Sigma\rangle\oplus(\mathbf{Z}^{d}\otimes\langle\Sigma\rangle\big{)}\Big{]}&\text{if }\eta\cdot\Sigma\in\Theta_{d+1}\text{is not divisible by }2\\ \hfil\mathrm{SL}_{d}(\mathbf{Z})\ltimes\Big{[}\Omega/\big{(}\mathbf{Z}^{d}\otimes\langle\Sigma\rangle\big{)}\Big{]}&\text{if }\eta\cdot\Sigma\in\Theta_{d+1}\text{ is divisible by }2\end{cases}

which follows by replacing Ω\Omega by Ω/(𝐙dΣ)\Omega/(\mathbf{Z}^{d}\otimes\langle\Sigma\rangle) in the proof of 2.5.

3. Splitting the homology action and the proof of A

To deduce A from D, we first determine for which homotopy tori 𝒯\mathcal{T} the map π0Diff+(𝒯)SLd(𝐙)\pi_{0}\,\mathrm{Diff}^{+}(\mathcal{T})\rightarrow\mathrm{SL}_{d}(\mathbf{Z}) is surjective. The following was stated in [BT21, p. 4] without proof.

Lemma 3.1.

For a homotopy torus 𝒯\mathcal{T} of dimension d4d\neq 4, the map π0Diff+(𝒯)SLd(𝐙)\pi_{0}\,\mathrm{Diff}^{+}(\mathcal{T})\to\mathrm{SL}_{d}(\mathbf{Z}) is surjective if and only if 𝒯\mathcal{T} is diffeomorphic to Td#ΣT^{d}\#\Sigma for some ΣΘd\Sigma\in\Theta_{d}.

Proof.

The direction \Leftarrow is easy: if 𝒯TdΣ\mathcal{T}\cong T^{d}\sharp\Sigma, then π0Diff+(𝒯)SLd(𝐙)\pi_{0}\,\mathrm{Diff}^{+}(\mathcal{T})\to\mathrm{SL}_{d}(\mathbf{Z}) is surjective because we can precompose it with the map ext:π0Diff((Td))π0Diff+(𝒯)\mathrm{ext}_{*}\colon\pi_{0}\,\mathrm{Diff}_{\partial}((T^{d})^{\circ})\rightarrow\pi_{0}\,\mathrm{Diff}^{+}(\mathcal{T}) and use that π0Diff((Td))SLd(𝐙)\pi_{0}\,\mathrm{Diff}_{\partial}((T^{d})^{\circ})\rightarrow\mathrm{SL}_{d}(\mathbf{Z}) is surjective which holds for instance as a result of 2.2.

For the direction \Rightarrow, we may assume d5d\geq 5 since for d3d\leq 3 any torus 𝒯\mathcal{T} is diffeomorphic to the standard torus TdT^{d}. This allows us to use smoothing theory [KS77, Essay V] which we briefly recall in a form suitable for our purposes: given a closed topological manifold MM of dimension d5d\geq 5, the set Smcon(M)\mathrm{Sm}^{\mathrm{con}}(M) of concordance classes of smooth structures on MM is the set of equivalence classes of pairs (T,ψ)(T,\psi) of a smooth manifold TT together with a homeomorphism ψ:TM\psi\colon T\to M, where two pairs (T,ψ)(T,\psi) and (T,ψ)(T^{\prime},\psi^{\prime}) are equivalent if there is a diffeomorphism Φ:TT\Phi\colon T\rightarrow T^{\prime} such that the homeomorphisms ψ\psi and ψΦ\psi^{\prime}\circ\Phi are concordant. The group π0Homeo~(M)\smash{\pi_{0}\,\widetilde{\mathrm{Homeo}}(M)} of concordance classes of homeomorphisms acts on Smcon(M)\mathrm{Sm}^{\mathrm{con}}(M) by postcomposition and the set of orbits is in bijection with the set Smdiff(M)\mathrm{Sm}^{\mathrm{diff}}(M) of diffeomorphism classes of smooth manifolds homeomorphic to MM, induced by sending (T,ψ)(T,\psi) to TT. There is a map

η:Smcon(M)Lift(M,BOBTop)\eta\colon\mathrm{Sm}^{\mathrm{con}}(M)\longrightarrow\mathrm{Lift}(M,\mathrm{BO}\rightarrow\mathrm{BTop})

to the set of isomorphism classes of pairs of a stable vector bundle over MM together with an isomorphism of the underlying stable Euclidean bundle with the stable topological tangent bundle τMTop\tau^{\mathrm{Top}}_{M} of MM. The map η\eta is given by assigning a pair (T,ψ)(T,\psi) to the pullback (ψ1)τTDiff(\psi^{-1})^{*}\tau^{\mathrm{Diff}}_{T} of the stable tangent bundle of TT along ψ1\psi^{-1}, together with the isomorphism induced by the stable topological derivative of ψ1\psi^{-1}. The map η\eta turns out to be a bijection, by one of the main results of smoothing theory. Unwrapping definitions, one sees that the action of απ0Homeo~(M)\smash{\alpha\in\pi_{0}\,\widetilde{\mathrm{Homeo}}(M)} on [T,ψ]Smcon(M)[T,\psi]\in\mathrm{Sm}^{\mathrm{con}}(M) is induced by pulling back the bundle along ψ1\psi^{-1} and postcomposing the isomorphism with the stable topological derivative of ψ1\psi^{-1}. The set Lift(M,BOBTop)\mathrm{Lift}(M,\mathrm{BO}\rightarrow\mathrm{BTop}) is a torsor for the group [M,Top/O][M,\mathrm{Top}/\mathrm{O}] of stable vector bundles on MM together with a trivialisation of the underlying stable Euclidean bundle; the group structure and the action are induced by taking direct sums. Thus, if MM comes already equipped with a smooth structure then we obtain a bijection Lift(M,BOBTop)[M,Top/O]\mathrm{Lift}(M,\mathrm{BO}\rightarrow\mathrm{BTop})\cong[M,\mathrm{Top}/\mathrm{O}], postcomposition with which gives a bijection

δ:Smcon(M)[M,Top/O].\delta\colon\mathrm{Sm}^{\mathrm{con}}(M)\longrightarrow[M,\mathrm{Top}/\mathrm{O}].

Going through the definition, the action of απ0Homeo~(M)\smash{\alpha\in\pi_{0}\,\widetilde{\mathrm{Homeo}}(M)} on [T,ψ]Sm(M)[T,\psi]\in\mathrm{Sm}(M) translates to δ(T,αψ)=(α1)δ(T,ψ)+δ(M,α)\delta(T,\alpha\circ\psi)=(\alpha^{-1})^{*}\delta(T,\psi)+\delta(M,\alpha).

We now specialise to M=TdM=T^{d}. Given a homotopy torus 𝒯\mathcal{T} of dimension d5d\geq 5, a choice of homeomorphism φ:𝒯Td\varphi\colon\mathcal{T}\to T^{d} induces a class [𝒯,φ]Smcon(Td)[\mathcal{T},\varphi]\in\mathrm{Sm}^{\mathrm{con}}(T^{d}) and a morphism π0Diff+(𝒯)π0Homeo~+(Td)\pi_{0}\,\mathrm{Diff}^{+}(\mathcal{T})\rightarrow\smash{\pi_{0}\,\widetilde{\mathrm{Homeo}}^{+}(T^{d})} by conjugation with φ\varphi. This agrees with the map to SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) when precomposed with the action map π0Homeo~+(Td)SLd(𝐙)\smash{\pi_{0}\,\widetilde{\mathrm{Homeo}}^{+}(T^{d})}\rightarrow\mathrm{SL}_{d}(\mathbf{Z}). The latter is an isomorphism as a result of 1.11 and the isomorphism π0Homeo~+(Td)π0Homeo~+(Td)\smash{\pi_{0}\,\widetilde{\mathrm{Homeo}}^{+}_{\ast}(T^{d})\cong\pi_{0}\,\widetilde{\mathrm{Homeo}}^{+}(T^{d})} (use the action of TdT^{d} on itself to see this), so it suffices to show that π0Diff+(𝒯)π0Homeo~+(Td)\pi_{0}\,\mathrm{Diff}^{+}(\mathcal{T})\rightarrow\smash{\pi_{0}\,\widetilde{\mathrm{Homeo}}^{+}(T^{d})} is not surjective unless 𝒯\mathcal{T} is diffeomorphic to TdΣT^{d}\sharp\Sigma for some ΣΘd\Sigma\in\Theta_{d}. The image of π0Diff+(𝒯)π0Homeo~(Td)\pi_{0}\,\mathrm{Diff}^{+}(\mathcal{T})\rightarrow\smash{\pi_{0}\,\widetilde{\mathrm{Homeo}}(T^{d})} is contained in the stabiliser of [𝒯,φ]Smcon(Td)[\mathcal{T},\varphi]\in\mathrm{Sm}^{\mathrm{con}}(T^{d}), so it is enough to show that [𝒯,φ]Sm(Td)[\mathcal{T},\varphi]\in\mathrm{Sm}(T^{d}) is not contained in the invariants of this action unless 𝒯TdΣ\mathcal{T}\cong T^{d}\sharp\Sigma for some ΣΘd\Sigma\in\Theta_{d}. Since any απ0Homeo~+(Td)SLd(𝐙)\smash{\alpha\in\pi_{0}\,\widetilde{\mathrm{Homeo}}^{+}(T^{d})\cong\mathrm{SL}_{d}(\mathbf{Z})} is isotopic to a diffeomorphism of TdT^{d}, the terms δ(Td,α)\delta(T^{d},\alpha) in the above description of the π0Homeo~(Td)\pi_{0}\,\widetilde{\mathrm{Homeo}}(T^{d})-action Smcon(Td)[Td,Top/O]\mathrm{Sm}^{\mathrm{con}}(T^{d})\cong[T^{d},\mathrm{Top}/\mathrm{O}] vanishes, and thus the action is simply by precomposition. In particular, it is an action by group homomorphisms if we equip [Td,Top/O][T^{d},\mathrm{Top}/\mathrm{O}] with the group structure induced by the infinite loop space structure on Top/O\mathrm{Top}/\mathrm{O}. Using this infinite loop space structure and the fact that TdT^{d} stably splits into a wedge of spheres we also get a direct sum decomposition of SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z})-modules [Td,Top/O]r=1dHom(Λr𝐙d,πrTop/O)\smash{[T^{d},\mathrm{Top}/\mathrm{O}]\cong\oplus_{r=1}^{d}\mathrm{Hom}(\Lambda^{r}\mathbf{Z}^{d},\pi_{r}\,\mathrm{Top}/\mathrm{O})}. We will show below that the invariants of this action are given by the subgroup Hom(Λd𝐙d,πdTop/O)πdTop/OΘd\mathrm{Hom}(\Lambda^{d}\mathbf{Z}^{d},\pi_{d}\,\mathrm{Top}/\mathrm{O})\cong\pi_{d}\,\mathrm{Top}/\mathrm{O}\cong\Theta_{d}. This will imply the claim, since the subgroup Θd[Td,Top/O]Smcon(Td)\Theta_{d}\leq[T^{d},\mathrm{Top}/\mathrm{O}]\cong\mathrm{Sm}^{\mathrm{con}}(T^{d}) corresponds to the classes of the pairs (TdΣ,idTdβ)(T^{d}\sharp\Sigma,\mathrm{id}_{T^{d}}\sharp\beta) where β:ΣSd\beta\colon\Sigma\rightarrow S^{d} is the unique homeomorphism up to isotopy that fixes the disc where the connected sum is taken, so in particular [𝒯,φ]Smcon(Td)[\mathcal{T},\varphi]\in\mathrm{Sm}^{\mathrm{con}}(T^{d}) is not contained in this subgroup unless 𝒯\mathcal{T} is diffeomorphic to TdΣT^{d}\sharp\Sigma for some ΣΘd\Sigma\in\Theta_{d}.

To finish the proof, it thus suffices to show that for a finitely generated abelian group AA, the SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z})-action on Hom(Λr𝐙d,A)\mathrm{Hom}(\Lambda^{r}\mathbf{Z}^{d},A) by precomposition with the inverse has no invariants for 0<r<d0<r<d. This is isomorphic to the standard action on Λr𝐙dA\Lambda^{r}\mathbf{Z}^{d}\otimes A up to the automorphism of SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) given by taking inverse transpose, so we may equivalently show that Λr𝐙dA\Lambda^{r}\mathbf{Z}^{d}\otimes A has no invariants . Without loss of generality we may assume that A=𝐙/nA=\mathbf{Z}/n is cyclic. In this case, Λr𝐙d𝐙/n\Lambda^{r}\mathbf{Z}^{d}\otimes\mathbf{Z}/n has a basis as a 𝐙/n\mathbf{Z}/n-module indexed by subsets I{1,,d}I\subset\{1,\ldots,d\} of cardinality rr, where the basis vector xIx_{I} corresponding to I{1,,d}I\subset\{1,\ldots,d\} is xi1xirx_{i_{1}}\wedge\cdots\wedge x_{i_{r}} for i1<<iri_{1}<\ldots<i_{r} and I={i1,,ir}I=\{i_{1},\ldots,i_{r}\} where x1,xdx_{1},\ldots x_{d} is the standard 𝐙/n\mathbf{Z}/n-basis of 𝐙d𝐙/n\mathbf{Z}^{d}\otimes\mathbf{Z}/n. Now observe that an elementary matrix (I+Eij)SLd(𝐙)(I+E_{ij})\in\mathrm{SL}_{d}(\mathbf{Z}) for 1i,jd1\leq i,j\leq d acts by sending xIx_{I} to xI±x(I\i)jx_{I}\pm x_{(I\backslash i)\cup j} if iIi\in I and jIj\notin I, and to itself otherwise. On a general element v=IλI(v)xIΛr𝐙d𝐙/n\smash{v=\sum_{I}\lambda_{I}(v)\cdot x_{I}\in\Lambda^{r}\mathbf{Z}^{d}\otimes\mathbf{Z}/n}, the matrix (I+Eij)(I+E_{ij}) thus acts by

viI or jIλI(v)xI+iI and jI(λI(v)+λ(I\j)i(v))xI,\textstyle{v\longmapsto\underset{i\in I\text{ or }j\not\in I}{\sum}\lambda_{I}(v)\cdot x_{I}+\underset{i\not\in I\text{ and }j\in I}{\sum}(\lambda_{I}(v)+\lambda_{(I\backslash j)\cup i}(v))\cdot x_{I}},

so if vv is an invariant, then λ(I\j)i(v)=0\lambda_{(I\backslash j)\cup i}(v)=0 for all II with iIi\not\in I and jIj\in I. But since 1i,jd1\leq i,j\leq d were arbitrary and 0<r<d0<r<d, these are in fact all coefficients, so all invariants are zero. ∎

3.1. Proof of A

We conclude this section with the proof of A, which says that the map π0Diff+(𝒯)SLd(𝐙)\pi_{0}\,\mathrm{Diff}^{+}(\mathcal{T})\to\mathrm{SL}_{d}(\mathbf{Z}) given by the action on H1(𝒯)=𝐙dH_{1}(\mathcal{T})=\mathbf{Z}^{d} admits a splitting if and only if 𝒯=TdΣ\mathcal{T}=T^{d}\sharp\Sigma for ΣΘd\Sigma\in\Theta_{d} such that ηΣΘd+1\eta\cdot\Sigma\in\Theta_{d+1} divisible by 2.

Proof of A.

We distinguish the cases whether a given homotopy torus 𝒯\mathcal{T} of dimension d4d\neq 4 is diffeomorphic to TdΣT^{d}\sharp\Sigma for some ΣΘd\Sigma\in\Theta_{d} or not. If it is not, then the map π0Diff+(𝒯)SLd(𝐙)\pi_{0}\,\mathrm{Diff}^{+}(\mathcal{T})\rightarrow\mathrm{SL}_{d}(\mathbf{Z}) is not surjective by 3.1, so it is in particular not split surjective. If it is, then by D the group π0Diff+(𝒯)\pi_{0}\,\mathrm{Diff}^{+}(\mathcal{T}) is isomorphic, compatibly with the map to SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}), to a semidirect product of SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) or SL¯d(𝐙)\overline{\mathrm{SL}}_{d}(\mathbf{Z}) depending on whether ηΣΘd+1\eta\cdot\Sigma\in\Theta_{d+1} is divisible by 22 or not. In the first case, the map to SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) visibly admits a splitting. In the second case, a hypothetical splitting would in particular induce a splitting of the projection SL¯d(𝐙)SLd(𝐙)\overline{\mathrm{SL}}_{d}(\mathbf{Z})\rightarrow\mathrm{SL}_{d}(\mathbf{Z}), which does not exist since this extension is nontrivial. This finishes the proof. ∎

4. Endomorphisms of SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) and the proofs of Theorems B and E

This section serves to deduce B from the classification result for endomorphisms of SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) stated as E, and to prove the latter.

4.1. Proof of B assuming E

Assuming E, we prove B. We first assume G=Homeo+(𝒯)Homeo+(Td)G=\mathrm{Homeo}^{+}(\mathcal{T})\cong\mathrm{Homeo}^{+}(T^{d}). Given a nontrivial homomorphism φ:SLd(𝐙)Homeo+(Td)\varphi\colon\mathrm{SL}_{d}(\mathbf{Z})\rightarrow\mathrm{Homeo}^{+}(T^{d}) for d3d\geq 3, the composition with the action on homology Homeo+(Td)SLd(𝐙)\mathrm{Homeo}^{+}(T^{d})\rightarrow\mathrm{SL}_{d}(\mathbf{Z}) is by E either trivial or an isomorphism, so we have to exclude the former. If it were trivial, then φ\varphi would have image in TorTop(Td)=ker(Homeo+(Td)SLd(𝐙))\mathrm{Tor}^{\mathrm{Top}}(T^{d})=\ker(\mathrm{Homeo}^{+}(T^{d})\rightarrow\mathrm{SL}_{d}(\mathbf{Z})). Suppose for contradiction that φ:SLd(𝐙)TorTop(Td)\varphi\colon\mathrm{SL}_{d}(\mathbf{Z})\rightarrow\mathrm{Tor}^{\mathrm{Top}}(T^{d}) is nontrivial. Its kernel is a normal subgroup, so by [Men65, Corollary 1, p. 36] it is either (a) contained in the centre Z(SLd(𝐙))Z(\mathrm{SL}_{d}(\mathbf{Z})), which is trivial or 𝐙/2\mathbf{Z}/2 depending on the parity of dd, or (b) of finite index. In either case, the image of ϕ\phi contains a nonabelian finite group HH: in case (a) it contains SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) or PSLd(𝐙)\mathrm{PSL}_{d}(\mathbf{Z}), so in particular a nonabelian finite group HH, and in case (b) the image of φ\varphi is finite itself, and also nonabelian since otherwise φ\varphi would be trivial since SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) is perfect for d3d\geq 3 (see 2.1).

To make use of the nonabelian finite subgroup HTorTop(Td)H\leq\mathrm{Tor}^{\mathrm{Top}}(T^{d}), following [LR81], we consider the extension 0𝐙dNHomeo(Td~)(𝐙d)Homeo(Td)00\rightarrow\mathbf{Z}^{d}\rightarrow N_{\mathrm{Homeo}(\smash{\widetilde{T^{d}}})}(\mathbf{Z}^{d})\rightarrow\mathrm{Homeo}(T^{d})\rightarrow 0 whose middle group is the normalizer of 𝐙d=π1(Td)\mathbf{Z}^{d}=\pi_{1}(T^{d}) considered as a subgroup of the homeomorphism group Homeo(Td~)\mathrm{Homeo}(\smash{\widetilde{T^{d}}}) of the universal cover. Note that the induced action of Homeo(Td)\mathrm{Homeo}(T^{d}) by 𝐙d\mathbf{Z}^{d} agrees by construction with the action on the fundamental group. The pullback 0𝐙dEH00\rightarrow\mathbf{Z}^{d}\rightarrow E\rightarrow H\rightarrow 0 of this extension along HHomeo(Td)H\leq\mathrm{Homeo}(T^{d}) is, by the Corollary on p. 256 of loc.cit. admissible in the sense of p. 256 loc.cit.. The proof of Proposition 2 loc.cit. then shows that the centraliser CE(𝐙d)C_{E}(\mathbf{Z}^{d}) of 𝐙d\mathbf{Z}^{d} in EE is abelian. But since HTorTop(Td)H\leq\mathrm{Tor}^{\mathrm{Top}}(T^{d}) acts trivially on 𝐙d\mathbf{Z}^{d} we have CE(𝐙d)=EC_{E}(\mathbf{Z}^{d})=E, so EE is abelian and thus the same holds for HH which cannot be true by the choice of HH, so φ\varphi has to be trivial.

The case G=Diff+(𝒯)G=\mathrm{Diff}^{+}(\mathcal{T}) follows from the case Homeo+(𝒯)\mathrm{Homeo}^{+}(\mathcal{T}) by postcomposing a given homomorphism φ:SLd(𝐙)Diff+(𝒯)\varphi\colon\mathrm{SL}_{d}(\mathbf{Z})\rightarrow\mathrm{Diff}^{+}(\mathcal{T}) with the inclusion Diff+(𝒯)Homeo+(𝒯)\mathrm{Diff}^{+}(\mathcal{T})\leq\mathrm{Homeo}^{+}(\mathcal{T}), so we are left to prove the addendum concerning homomorphisms from SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) into π0Diff+(𝒯)\pi_{0}\,\mathrm{Diff}^{+}(\mathcal{T}) or π0Homeo+(𝒯)\pi_{0}\,\mathrm{Homeo}^{+}(\mathcal{T}) under the additional assumption d4,5d\neq 4,5. By the same argument as before, it suffices to show that all homomorphisms from SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) into π0TorDiff(𝒯)\pi_{0}\,\mathrm{Tor}^{\mathrm{Diff}}(\mathcal{T}) or π0TorTop(𝒯)\pi_{0}\,\mathrm{Tor}^{\mathrm{Top}}(\mathcal{T}) are trivial. 4.1 below says that the latter two groups are abelian, so such morphisms factor over the (trivial) abelianisation of SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) (see 2.1) and are therefore trivial, as claimed.

Lemma 4.1.

For a homotopy torus 𝒯\mathcal{T} of dimension d4,5d\neq 4,5, the kernels of the homology actions

π0TorTop(𝒯)\displaystyle\pi_{0}\,\mathrm{Tor}^{\mathrm{Top}}(\mathcal{T}) =ker[π0Homeo+(𝒯)SLd(𝐙)]\displaystyle=\ker\big{[}\pi_{0}\,\mathrm{Homeo}^{+}(\mathcal{T})\rightarrow\mathrm{SL}_{d}(\mathbf{Z})\big{]}
π0TorDiff(𝒯)\displaystyle\pi_{0}\,\mathrm{Tor}^{\mathrm{Diff}}(\mathcal{T}) =ker[π0Diff+(𝒯)SLd(𝐙)]\displaystyle=\ker\big{[}\pi_{0}\,\mathrm{Diff}^{+}(\mathcal{T})\rightarrow\mathrm{SL}_{d}(\mathbf{Z})\big{]}

are both abelian.

Proof.

For d3d\leq 3, the homotopy torus 𝒯\mathcal{T} is diffeomorphic to the standard torus and both kernels π0TorTop(𝒯)\pi_{0}\,\mathrm{Tor}^{\mathrm{Top}}(\mathcal{T}) and π0TorDiff(𝒯)\pi_{0}\,\mathrm{Tor}^{\mathrm{Diff}}(\mathcal{T}) are trivial, so in particular abelian. To show the claim for d6d\geq 6, note that π0TorTop(𝒯)π0TorTop(Td)\pi_{0}\,\mathrm{Tor}^{\mathrm{Top}}(\mathcal{T})\cong\pi_{0}\,\mathrm{Tor}^{\mathrm{Top}}(T^{d}) because 𝒯\mathcal{T} is homeomorphic to TdT^{d}, so π0TorTop(𝒯)\pi_{0}\,\mathrm{Tor}^{\mathrm{Top}}(\mathcal{T}) is abelian since we have π0TorTop(Td)(𝐙/2)\pi_{0}\,\mathrm{Tor}^{\mathrm{Top}}(T^{d})\cong(\mathbf{Z}/2)^{\infty} by [Hat78, Theorem 4.1]. To show that π0TorDiff(𝒯)\pi_{0}\,\mathrm{Tor}^{\mathrm{Diff}}(\mathcal{T}) is abelian, note that as 𝒯K(𝐙d,1)\mathcal{T}\simeq K(\mathbf{Z}^{d},1) we may view the map π0Diff+(𝒯)SLd(𝐙)\pi_{0}\,\mathrm{Diff}^{+}(\mathcal{T})\rightarrow\mathrm{SL}_{d}(\mathbf{Z}) as the induced map on path components of the map Diff+(𝒯)hAut+(𝒯)\mathrm{Diff}^{+}(\mathcal{T})\rightarrow\mathrm{hAut}^{+}(\mathcal{T}) to the space of orientation-preserving homotopy equivalences, so π0TorDiff(𝒯)\pi_{0}\,\mathrm{Tor}^{\mathrm{Diff}}(\mathcal{T}) receives an epimorphism from π1(hAut+(𝒯)/Diff+(𝒯))\pi_{1}(\mathrm{hAut}^{+}(\mathcal{T})/\mathrm{Diff}^{+}(\mathcal{T})). Replacing TdT^{d} by 𝒯\mathcal{T} in the argument for (3) on page 8 of [Hat78] and using that 𝒯\mathcal{T} is homeomorphic to TdT^{d}, we get that π1(hAut+(𝒯)/Diff+(𝒯))\pi_{1}(\mathrm{hAut}^{+}(\mathcal{T})/\mathrm{Diff}^{+}(\mathcal{T})) is isomorphic to the abelian group (0jd(Λj𝐙d)Θdj+1)((Λd2𝐙d)𝐙/2)𝐙/2(\oplus_{0\leq j\leq d}(\Lambda^{j}\mathbf{Z}^{d})\otimes\Theta_{d-j+1})\oplus((\Lambda^{d-2}\mathbf{Z}^{d})\otimes\mathbf{Z}/2)\oplus\mathbf{Z}/2^{\infty}, so the claim follows (the final step can also be proved via smoothing theory). ∎

4.2. Proof of E

In the remainder of this section, we prove E. The proof makes use of the subgroup 𝐔d<SLd(𝐙)\mathbf{U}_{d}<\mathrm{SL}_{d}(\mathbf{Z}) of unipotent upper triangular matrices which in particular contains the elementary matrices EijE_{ij} for 1i<jd1\leq i<j\leq d; these have 11 on the diagonal and at the (i,j)(i,j)th entry, and 0 at all other entries. It is well-known that 𝐔d\mathbf{U}_{d} is an (d1)(d-1)-step nilpotent group whose centre is generated by the elementary matrix E1dE_{1d}, which is an iterated commutator of length (d1)(d-1), namely E1d=[E12,[E23,[,[Ed2,d1,Ed1,d]]]]E_{1d}=[E_{12},[E_{23},[\ldots,[E_{d-2,d-1},E_{d-1,d}]]]]. An important ingredient in the proof of E is the following lemma on complex representations of 𝐔d\mathbf{U}_{d}. In its statement and in all that follows, we write

()t:SLd(𝐙)SLd(𝐙)(-)^{-t}\colon\mathrm{SL}_{d}(\mathbf{Z})\longrightarrow\mathrm{SL}_{d}(\mathbf{Z})

for the automorphism of SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) given by taking inverse-transpose.

Lemma 4.2.

Fix d3d\geq 3 and a homomorphism ϕ:𝐔dGLm(𝐂)\phi\colon\mathbf{U}_{d}\rightarrow\mathrm{GL}_{m}(\mathbf{C}) with mdm\leq d.

  1. (i)

    Assume d4d\geq 4. If m<dm<d, or if m=dm=d and ϕ(E1d)\phi(E_{1d}) is not a scalar, then ϕ(E1d)\phi(E_{1d}) is unipotent.

  2. (ii)

    If m<dm<d and ϕ(Eij)\phi(E_{ij}) is unipotent for each i<ji<j, then ϕ(E1d)=id\phi(E_{1d})=\mathrm{id}.

  3. (iii)

    If m=dm=d and ϕ(Eij)\phi(E_{ij}) is unipotent for each i<ji<j and ϕ(E1d)id\phi(E_{1d})\neq\mathrm{id}, then ϕ(E1d)id\phi(E_{1d})-\mathrm{id} has rank 11.

  4. (iv)

    If ϕ(Eij)\phi(E_{ij}) is unipotent and ϕ(Eij)id\phi(E_{ij})-\mathrm{id} has rank 11 for all i<ji<j, then after possibly precomposing ϕ\phi with ()t(-)^{-t}, the matrices ϕ(E1d),,ϕ(Ed1d)\phi(E_{1d}),\ldots,\phi(E_{d-1d}) all have the same fixed set.

Remark 4.3.

The argument in Bader’s MathOverflow post [Mat17] contains the claim that for any representation ϕ:𝐔3GL3(𝐂)\phi\colon\mathbf{U}_{3}\rightarrow\mathrm{GL}_{3}(\mathbf{C}) the matrix ϕ(E13)\phi(E_{13}) is unipotent. This is incorrect: 4.4 (iii) gives a representation 𝐔3GL3(𝐂)\mathbf{U}_{3}\rightarrow\mathrm{GL}_{3}(\mathbf{C}) for which ϕ(E13)\phi(E_{13}) is a nontrivial scalar. Also 4.2 (i) fails for d=m=3d=m=3 (the case (ii) of 4.4 below involves representations 𝐔3GL3(𝐂)\mathbf{U}_{3}\rightarrow\mathrm{GL}_{3}(\mathbf{C}) for which ϕ(E13)\phi(E_{13}) is a non-scalar semisimple matrix).

For d=3d=3, we will circumvent the use of 4.2 (i) in later proofs by means of the following:

Lemma 4.4.

Fix a homomorphism ϕ:𝐔3GLm(𝐂)\phi\colon\mathbf{U}_{3}\rightarrow\mathrm{GL}_{m}(\mathbf{C}). If m=2m=2, then either

  1. (i)

    ϕ(E13)\phi(E_{13}) is unipotent, or

  2. (ii)

    there are μ,ν𝐂×\mu,\nu\in\mathbf{C}^{\times} and CGL2(𝐂)C\in\mathrm{GL}_{2}(\mathbf{C}) so that after postcomposing ϕ\phi with conjugation by CC,

    E12(μ00μ)E23(0ν10)E13(1001).E_{12}\mapsto\left(\begin{array}[]{ccc}\mu&0\\ 0&-\mu\end{array}\right)\>\>\>\>\>\>E_{23}\mapsto\left(\begin{array}[]{ccc}0&\nu\\ 1&0\\ \end{array}\right)\>\>\>\>\>\>E_{13}\mapsto\left(\begin{array}[]{ccc}-1&0\\ 0&-1\\ \end{array}\right).

If m=3m=3, then either

  1. (i)

    ϕ(E13)\phi(E_{13}) is unipotent,

  2. (ii)

    ϕ=ϕ1ϕ2\phi=\phi_{1}\oplus\phi_{2}, where ϕi:𝐔3GLi(𝐂)\phi_{i}\colon\mathbf{U}_{3}\rightarrow\mathrm{GL}_{i}(\mathbf{C}), up to conjugation, or

  3. (iii)

    there are λ,μ,ν𝐂×\lambda,\mu,\nu\in\mathbf{C}^{\times} and CGL3(𝐂)C\in\mathrm{GL}_{3}(\mathbf{C}) so that λ\lambda is a nontrivial cube root of 11, and after postcomposing ϕ\phi with conjugation by CC,

    E12(μ000λμ000λ2μ)E23(00ν100010)E13(λ000λ000λ).E_{12}\mapsto\left(\begin{array}[]{ccc}\mu&0&0\\ 0&\lambda\mu&0\\ 0&0&\lambda^{2}\mu\end{array}\right)\>\>\>\>\>\>E_{23}\mapsto\left(\begin{array}[]{ccc}0&0&\nu\\ 1&0&0\\ 0&1&0\end{array}\right)\>\>\>\>\>\>E_{13}\mapsto\left(\begin{array}[]{ccc}\lambda&0&0\\ 0&\lambda&0\\ 0&0&\lambda\end{array}\right).

We omit the proof of 4.4 since it is based on similar (and easier) analysis as the base case in the proof of 4.2 (i) which we explain now.

Proof of Lemma 4.2 (i).

We do an induction on dd. To simplify the notation we set uij:-ϕ(Eij)u_{ij}\coloneq\phi(E_{ij}).

Base case. We treat the case d=4d=4 by hand. To show that u14u_{14} is unipotent, it suffices to prove that all its eigenvalues λ\lambda equal 11. Let VλV_{\lambda} be the λ\lambda-eigenspace for u14u_{14}. Since E14E_{14} is central in 𝐔4\mathbf{U}_{4}, restricting to VλV_{\lambda} gives a homomorphism 𝐔4GL(Vλ)\mathbf{U}_{4}\rightarrow\mathrm{GL}(V_{\lambda}) whose image of EijE_{ij} we denote by uiju_{ij}^{\prime}. Next we distinguish cases depending on the dimension of VλV_{\lambda}. By the assumption that u14=ϕ(E14)u_{14}=\phi(E_{14}) is not a scalar when m=dm=d, we know dim(Vλ)3\dim(V_{\lambda})\leq 3. If dimVλ=1\dim V_{\lambda}=1, then since GL1(𝐂)=𝐂×\mathrm{GL}_{1}(\mathbf{C})=\mathbf{C}^{\times} is abelian, we have u14=1u_{14}^{\prime}=1 because E14=[E13,E34]E_{14}=[E_{13},E_{34}] is a commutator, so λ=1\lambda=1. If dimVλ=2\dim V_{\lambda}=2, we consider the subgroup u13,u34,u14GL(Vλ)\langle u_{13}^{\prime},u_{34}^{\prime},u_{14}^{\prime}\rangle\leq\mathrm{GL}(V_{\lambda}) generated by the images of E13E_{13}, E34E_{34}, and E14E_{14} in GL(Vλ)\mathrm{GL}(V_{\lambda}). By assumption u14=λid2×2u_{14}^{\prime}=\lambda\cdot\mathrm{id}_{2\times 2}. Let xVλx\in V_{\lambda} be an eigenvector for u13u_{13}^{\prime} with eigenvalue μ\mu. Using the relation [E13,E34]=E14[E_{13},E_{34}]=E_{14} we conclude that (u34)i(x)(u_{34}^{\prime})^{i}(x) is an eigenvector for u13u_{13}^{\prime} with eigenvalue λiμ\lambda^{i}\mu. Since dimVλ=2\dim V_{\lambda}=2, this forces λ2=1\lambda^{2}=1 because eigenvectors with different eigenvalues are linearly independent, and thus λ=±1\lambda=\pm 1. Suppose for a contradiction that λ=1\lambda=-1. Then u13u_{13}^{\prime} has two distinct eigenvalues μ\mu and μ-\mu. Since E13E_{13} is central in E12,E23𝐔3\langle E_{12},E_{23}\rangle\cong\mathbf{U}_{3}, we deduce that u12u_{12}^{\prime} and u23u_{23}^{\prime} are simultaneously diagonalisable; in particular they commute. But since E13=[E12,E23]E_{13}=[E_{12},E_{23}], this implies u13=idu_{13}^{\prime}=\mathrm{id}, which is a contradiction, so λ\lambda has to be 11. Finally, suppose that dimVλ=3\dim V_{\lambda}=3. In this case the argument is very similar to the preceding case: by assumption u14=λid3×3u_{14}^{\prime}=\lambda\cdot\mathrm{id}_{3\times 3}, and the relation [E13,E34]=E14[E_{13},E_{34}]=E_{14} implies that λ𝐂×\langle\lambda\rangle\subset\mathbf{C}^{\times} acts freely on the eigenvalues of u13u_{13}^{\prime} which implies λ3=1\lambda^{3}=1. If λ1\lambda\neq 1, then u13u_{13}^{\prime} has distinct eigenvalues μ\mu, λμ\lambda\mu, and λ2μ\lambda^{2}\mu for some μ\mu. Using the fact that E13E_{13} is both central and a commutator in E12,E23𝐔4\langle E_{12},E_{23}\rangle\subset\mathbf{U}_{4}, we reach a contradiction.

Induction step. Fix an eigenvalue λ\lambda for u1d=ϕ(E1d)u_{1d}=\phi(E_{1d}), and let VλV_{\lambda} be the corresponding eigenspace. We have to show λ=1\lambda=1. As in the base case, since u1du_{1d} is not a scalar if d=md=m, so dim(Vλ)d1\dim(V_{\lambda})\leq d-1 and since E1dE_{1d} is central in 𝐔d\mathbf{U}_{d}, the representation restricts to 𝐔dGL(Vλ)\mathbf{U}_{d}\rightarrow\mathrm{GL}(V_{\lambda}). As before we write uij\smash{u_{ij}^{\prime}} for the image of EijE_{ij} under this homomorphism. Consider the subgroup E1,d1,Ed1,d,E1,d𝐔3\langle E_{1,d-1},E_{d-1,d},E_{1,d}\rangle\cong\mathbf{U}_{3} of 𝐔d\mathbf{U}_{d}. Let μ\mu be an eigenvalue of u1,d1\smash{u_{1,d-1}^{\prime}} and let VμVλV_{\mu}\leq V_{\lambda} be the corresponding eigenspace. As above, λiμ\lambda^{i}\mu is also an eigenvalue for u1,d1\smash{u_{1,d-1}^{\prime}} for each ii. Consider the subgroup Eiji<jd1𝐔d1\langle E_{ij}\mid i<j\leq d-1\rangle\cong\mathbf{U}_{d-1} of 𝐔d\mathbf{U}_{d}. Since E1,d1E_{1,d-1} is central in this copy of 𝐔d1\mathbf{U}_{d-1}, there is an induced map 𝐔d1GL(Vμ)\mathbf{U}_{d-1}\rightarrow\mathrm{GL}(V_{\mu}), Eijuij′′E_{ij}\mapsto u_{ij}^{\prime\prime}. If λ1\lambda\neq 1, then VμV_{\mu} is a proper subspace of VλV_{\lambda} (since the eigenspaces for μ\mu and λμ\lambda\mu are linearly independent). Then dim(Vμ)d2\dim(V_{\mu})\leq d-2, so the induction hypothesis implies that u1,d1′′\smash{u_{1,d-1}^{\prime\prime}} is unipotent, so μ=1\mu=1. Since the same argument applies for each eigenspace of u1,d1u_{1,d-1}^{\prime}, we conclude that 1=μ=λμ1=\mu=\lambda\mu, so λ=1\lambda=1 as claimed. ∎

Proof of Lemma 4.2 (ii).

Fixing ϕ:𝐔dGLm(𝐂)\phi:\mathbf{U}_{d}\rightarrow\mathrm{GL}_{m}(\mathbf{C}) such that ϕ(Eij)\phi(E_{ij}) is unipotent for all 1i<jd1\leq i<j\leq d, we want to show ϕ(E1d)=id\phi(E_{1d})=\mathrm{id}. As before we write uij=ϕ(Eij)u_{ij}=\phi(E_{ij}). Note that the special case m=d1m=d-1 implies the case m<d1m<d-1, because if m<d1m<d-1 then we may restrict ϕ\phi to the subgroup 𝐔m+1Eij1i<jm+1𝐔d\mathbf{U}_{m+1}\cong\langle E_{ij}\mid 1\leq i<j\leq m+1\rangle\leq\mathbf{U}_{d} to conclude u1,m+1=idu_{1,m+1}=\mathrm{id} from the special case, so using E1d=[E1,m+1,Em+1,d]E_{1d}=[E_{1,m+1},E_{m+1,d}] we get u1d=[u1,m+1,um+1,d]=idu_{1d}=[u_{1,m+1},u_{m+1,d}]=\mathrm{id}. To prove the special case m=d1m=d-1, we do an induction on the dimension dd.

Base case. To settle the case ϕ:𝐔3GL2(𝐂)\phi\colon\mathbf{U}_{3}\rightarrow\mathrm{GL}_{2}(\mathbf{C}), suppose for a contradiction that u13u_{13} is not the identity. Since it is unipotent by assumption, it has up to conjugation the form (1101)\left(\begin{smallmatrix}1&1\\ 0&1\end{smallmatrix}\right), so by postcomposing ϕ\phi with this conjugation we may assume that u13u_{13} equals this matrix. Since E13E_{13} is central in 𝐔3\mathbf{U}_{3}, the image of ϕ\phi is contained in the centraliser of (1101)\left(\begin{smallmatrix}1&1\\ 0&1\end{smallmatrix}\right) which consists of matrices of the form (ab0a)\left(\begin{smallmatrix}a&b\\ 0&a\end{smallmatrix}\right). This is an abelian subgroup, so u13=[u12,u23]u_{13}=[u_{12},u_{23}] is identity, a contradiction.

Induction step. For the induction step, we fix ϕ:𝐔dGLd1(𝐂)\phi\colon\mathbf{U}_{d}\rightarrow\mathrm{GL}_{d-1}(\mathbf{C}) and suppose for a contradiction that u1didu_{1d}\neq\mathrm{id}. Consider the subspaces K1K2𝐂d1K_{1}\subset K_{2}\subset\mathbf{C}^{d-1} where Ki=ker(u1did)iK_{i}=\ker(u_{1d}-\mathrm{id})^{i}. Writing ki:-dimKik_{i}\coloneq\dim K_{i} we have k1>0k_{1}>0 since u1du_{1d} is unipotent, k2k1>0\ell\coloneqq k_{2}-k_{1}>0 since u1didu_{1d}\neq\mathrm{id}, and k1\ell\leq k_{1} (one way to see this is to consider the Jordan normal form). Note that since E1dE_{1d} is central in 𝐔d\mathbf{U}_{d}, the image of ϕ\phi preserves K2K_{2} so we obtain a morphism ϕ:𝐔dGL(K2)\phi^{\prime}\colon\mathbf{U}_{d}\rightarrow\mathrm{GL}(K_{2}) by restriction. We write uiju_{ij}^{\prime} for its image of EijE_{ij}. Setting m:-k10m\coloneq k_{1}-\ell\geq 0, we choose a basis for K2K_{2} that extends a basis for K1K_{1} and that has the property that

(16) u1d=(id×0id×0idm×m000id×.)u_{1d}^{\prime}=\left(\begin{array}[]{ccc}\mathrm{id}_{\ell\times\ell}&0&\mathrm{id}_{\ell\times\ell}\\ 0&\mathrm{id}_{m\times m}&0\\ 0&0&\mathrm{id}_{\ell\times\ell}.\end{array}\right)

in this basis. To see that such a basis exists, it is again helpful to use the Jordan normal form. Since E1dE_{1d} is central in 𝐔d\mathbf{U}_{d}, the morphism 𝐔dGL(K2)GL2+m(𝐂)\mathbf{U}_{d}\rightarrow\mathrm{GL}(K_{2})\cong\mathrm{GL}_{2\ell+m}(\mathbf{C}) lands in the centraliser of (16) which are the matrices of the form

(17) (AXZ0BY00A).\left(\begin{array}[]{ccc}A&X&Z\\ 0&B&Y\\ 0&0&A\end{array}\right).

We claim that u1,+m+1\smash{u^{\prime}_{1,\ell+m+1}} and u+m+1,2+m+1\smash{u^{\prime}_{\ell+m+1,2\ell+m+1}} have the form

(18) u1,+m+1=(id×0Z0idm×m000id×)andu+m+1,2+m+1=(id×XZ0BY00id×)u^{\prime}_{1,\ell+m+1}=\left(\begin{array}[]{ccc}\mathrm{id}_{\ell\times\ell}&0&Z\\ 0&\mathrm{id}_{m\times m}&0\\ 0&0&\mathrm{id}_{\ell\times\ell}\end{array}\right)\quad\text{and}\quad u^{\prime}_{\ell+m+1,2\ell+m+1}=\left(\begin{array}[]{ccc}\mathrm{id}_{\ell\times\ell}&X&Z^{\prime}\\ 0&B&Y\\ 0&0&\mathrm{id}_{\ell\times\ell}\end{array}\right)

for some B,X,Y,ZB,X,Y,Z, and ZZ^{\prime}. Assuming this claim for now, we observe that the matrices (18) commute, so u1,2+m+1=[u1,+m+1,u+m+1,2+m+1]\smash{u^{\prime}_{1,2\ell+m+1}=[u^{\prime}_{1,\ell+m+1},u^{\prime}_{\ell+m+1,2\ell+m+1}]} is the identity. If 2+m+1=d2\ell+m+1=d then we are done since this contradicts (16). If 2+m+1<d2\ell+m+1<d then the relation u1d=[u1,2+m+1,u2+m+1,d]\smash{u_{1d}^{\prime}=[u^{\prime}_{1,2\ell+m+1},u^{\prime}_{2\ell+m+1,d}]} shows that u1du_{1d}^{\prime} is the identity, which again contradicts (16). This leaves us with showing (18). We first treat u1,+m+1\smash{u^{\prime}_{1,\ell+m+1}}. Since ϕ\phi^{\prime} has image in (17), we may postcompose it with

(AXZ0BY00A)(AX0B) and (AXZ0BY00A)(BY0A).\left(\begin{array}[]{ccc}A&X&Z\\ 0&B&Y\\ 0&0&A\end{array}\right)\mapsto\left(\begin{array}[]{cc}A&X\\ 0&B\end{array}\right)\>\>\>\text{ and }\>\>\>\left(\begin{array}[]{ccc}A&X&Z\\ 0&B&Y\\ 0&0&A\end{array}\right)\mapsto\left(\begin{array}[]{cc}B&Y\\ 0&A\end{array}\right).

to obtain two homomorphisms 𝐔dGL+m(𝐂)\mathbf{U}_{d}\rightarrow\mathrm{GL}_{\ell+m}(\mathbf{C}). We may apply the induction hypothesis to the restriction of these to the subgroup 𝐔+m+1E1,i1<i+m+1\mathbf{U}_{\ell+m+1}\cong\langle E_{1,i}\mid 1<i\leq\ell+m+1\rangle to conclude that the image of u1,+m+1\smash{u^{\prime}_{1,\ell+m+1}} under these two homomorphism is the identity, so u1,+m+1\smash{u^{\prime}_{1,\ell+m+1}} has the claimed form.

To deal with the second matrix u+m+1,2+m+1\smash{u^{\prime}_{\ell+m+1,2\ell+m+1}} we argue similarly: postcompose ϕ\phi^{\prime} with the restriction to AA to obtain a morphism 𝐔dGL(𝐂)\mathbf{U}_{d}\rightarrow\mathrm{GL}_{\ell}(\mathbf{C}), restrict them to the subgroup 𝐔+1E+m+1,i+m+1<i2+m+1\mathbf{U}_{\ell+1}\cong\langle E_{\ell+m+1,i}\mid\ell+m+1<i\leq 2\ell+m+1\rangle in 𝐔d\mathbf{U}_{d}, and apply the induction hypothesis. ∎

Proof of 4.2 (iii).

Fix ϕ:𝐔dGLd(𝐂)\phi\colon\mathbf{U}_{d}\rightarrow\mathrm{GL}_{d}(\mathbf{C}) such that u1d=ϕ(E1d)u_{1d}=\phi(E_{1d}) is unipotent and nontrivial. The subspace K2=ker(u1did)2K_{2}=\ker(u_{1d}-\mathrm{id})^{2} is nontrivial, preserved by the image of ϕ\phi, each ϕ(Eij)\phi(E_{ij}) acts on it by a nontrivial unipotent, and ϕ(E1d)\phi(E_{1d}) acts nontrivially on it, so 4.2 (ii) implies K2=𝐂dK_{2}=\mathbf{C}^{d}. Arguing as in the proof of 4.2 (ii), up to changing basis (corresponding to postcomposing ϕ\phi with a conjugation), we can assume that (16) holds and by the same argument as in the previous proof ϕ\phi has image in matrices of the form (17) and u1,+m+1u_{1,\ell+m+1} has the form (18). We are left to show =1\ell=1 since then u1du_{1d} has rank 11 in view of (16). Assuming for a contradiction that >1\ell>1, then +m+1<d\ell+m+1<d, so u1d=[u1,+m+1,u+m+1,d]u_{1d}=[u_{1,\ell+m+1},u_{\ell+m+1,d}]. Written out in matrices this equation reads as

(id0id0id000id)=(id0Z0id000id)(AXZ0BY00A)(id0Z0id000id)(AXZ0BY00A)1\left(\begin{array}[]{ccc}\mathrm{id}&0&\mathrm{id}\\ 0&\mathrm{id}&0\\ 0&0&\mathrm{id}\end{array}\right)=\left(\begin{array}[]{ccc}\mathrm{id}&0&Z\\ 0&\mathrm{id}&0\\ 0&0&\mathrm{id}\end{array}\right)\left(\begin{array}[]{ccc}A&X&Z^{\prime}\\ 0&B&Y\\ 0&0&A\end{array}\right)\left(\begin{array}[]{ccc}\mathrm{id}&0&-Z\\ 0&\mathrm{id}&0\\ 0&0&\mathrm{id}\end{array}\right)\left(\begin{array}[]{ccc}A&X&Z^{\prime}\\ 0&B&Y\\ 0&0&A\end{array}\right)^{-1}

which implies ZAAZ=AZA-AZ=A, but this is a contradiction because the trace of ZAAZZA-AZ is 0, whereas that of AA is nonzero since AA is unipotent because so is u1du_{1d}^{\prime}, by assumption. ∎

Before proving Lemma 4.2 (iv), we discuss some properties of rank-1 operators. Given subspaces H,L𝐂dH,L\leq\mathbf{C}^{d} with dimH=d1\dim H=d-1, dimL=1\dim L=1, there is a rank-1 operator with kernel HH and image LL, which is unique up to a unit, namely the composition 𝐂d𝐂d/H𝐂𝐂L𝐂d\mathbf{C}^{d}\twoheadrightarrow\mathbf{C}^{d}/H\cong\mathbf{C}\rightarrow\mathbf{C}\cong L\hookrightarrow\mathbf{C}^{d}. In what follows, it will be convenient to consider rank-1 operators up to scalars; abusing notation, we will use ΠH,L\Pi_{H,L} to denote either this equivalence class of rank-1 operator with kernel HH and image LL. In terms of equivalence classes, the composition behaves as

ΠH,LΠH,L={0 if LH,ΠH,L if LH.\Pi_{H,L}\circ\Pi_{H^{\prime},L^{\prime}}=\begin{cases}0&\text{ if }L^{\prime}\subset H,\\ \Pi_{H^{\prime},L}&\text{ if }L^{\prime}\not\subset H.\end{cases}

The operator uH,Lid+ΠH,Lu_{H,L}\coloneqq\mathrm{id}+\Pi_{H,L} (which is well-defined up to scaling uH,Lidu_{H,L}-\mathrm{id} by a unit) is unipotent if and only if LHL\subset H (otherwise uH,Lu_{H,L} is diagonalisable and nontrivial). In this case the fixed set of uH,Lu_{H,L} is HH and its inverse is idΠH,L\mathrm{id}-\Pi_{H,L} which is another representative of uH,Lu_{H,L}. Fixing two such equivalence classes of unipotent operators uH,Lu_{H,L} and uH,Lu_{H^{\prime},L^{\prime}}, we have the commutator relation

(19) [uH,L,uH,L]={uH,L if LH and LH,uH,L if LHand LH,id if LH and LH.[u_{H,L}\>,\>u_{H^{\prime},L^{\prime}}]=\begin{cases}u_{H^{\prime},L}&\text{ if }L\subset H^{\prime}\text{ and }L^{\prime}\not\subset H,\\ u_{H,L^{\prime}}&\text{ if }L\not\subset H^{\prime}\,\,\text{and }L^{\prime}\subset H,\\ \mathrm{id}&\text{ if }L\subset H^{\prime}\ \,\text{ and }L^{\prime}\subset H.\end{cases}

If LHL\not\subset H^{\prime} and LHL^{\prime}\not\subset H, then the commutator is not unipotent.

The following observation will play a role in the proof of 4.2 (iv): Fixing unipotent operators uHi,Liu_{H_{i},L_{i}} as above for i=1,2,3i=1,2,3 and assuming firstly that uH1,L1u_{H_{1},L_{1}} commutes with uHj,Lju_{H_{j},L_{j}} for j=2,3j=2,3 and secondly that uH1,L1=[uH2,L2,uH3,L3]u_{H_{1},L_{1}}=[u_{H_{2},L_{2}}\>,\>u_{H_{3},L_{3}}], we may use the commutator formula from above to conclude that LjH1L_{j}\subset H_{1} for j=2,3j=2,3 and that H2=H1H_{2}=H_{1} or H3=H1H_{3}=H_{1}.

Proof of Lemma 4.2 (iv).

Since ϕ(Eij)id\phi(E_{ij})-\mathrm{id} has rank 11 for i<ji<j, the operators uij:-ϕ(Eij)=id+(ϕ(Eij)id)u_{ij}\coloneq\phi(E_{ij})=\mathrm{id}+(\phi(E_{ij})-\mathrm{id}) are for i<ji<j of the form uHij,Liju_{H_{ij},L_{ij}} as discussed above where HijH_{ij} is the kernel of ϕ(Eij)id\phi(E_{ij})-\mathrm{id}, i.e. the fixed set of ϕ(Eij)\phi(E_{ij}). We claim that either H1d=H2,d==Hd1,dH_{1d}=H_{2,d}=\cdots=H_{d-1,d} or H1d=H1,d1==H12H_{1d}=H_{1,d-1}=\cdots=H_{12}. This would imply the result, because the two cases are interchanged when precomposing ϕ\phi with ()t(-)^{-t}. To show this claim, we use that u1du_{1d} commutes with uiju_{ij} for i<ji<j. Since u1d=[u12,u2d]u_{1d}=[u_{12},u_{2d}], it follows from the discussion after (19) that either H12=H1dH_{12}=H_{1d} or H2d=H1dH_{2d}=H_{1d}. In the first case, we also have H1j=H1dH_{1j}=H_{1d} for all 2jd2\leq j\leq d, using u1j=[u12,u2j]u_{1j}=[u_{12},u_{2j}] and the fact that u2ju_{2j} preserves H1dH_{1d} since it commutes with u1du_{1d}. Similarly, in the second case we also have Hj,d=H1dH_{j,d}=H_{1d} for all 2jd2\leq j\leq d using uj,d=[uj,2,u2,d]u_{j,d}=[u_{j,2},u_{2,d}] and that uj,2u_{j,2} commutes with u1,du_{1,d}. ∎

We illustrate the utility of Lemma 4.2 to study representations of SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) by the following two corollaries, which will both play a role in the proof of Theorem E.

Corollary 4.5.

For d3d\geq 3 and m<dm<d, all homomorphisms ϕ:SLd(𝐙)GLm(𝐂)\phi\colon\mathrm{SL}_{d}(\mathbf{Z})\rightarrow\mathrm{GL}_{m}(\mathbf{C}) are trivial.

Under the additional assumption that ϕ\phi factors through SLm(𝐙)GLm(𝐂)\mathrm{SL}_{m}(\mathbf{Z})\leq\mathrm{GL}_{m}(\mathbf{C}), this corollary is proved in [Wei97, Lemma 3] using superrigidity and the congruence subgroup property.

Proof of 4.5.

If d4d\geq 4 then ϕ(E1d)\phi(E_{1d}) is unipotent by 4.2 (i) and since the EijE_{ij} are conjugate in SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) so are all ϕ(Eij)\phi(E_{ij}). We then apply 4.2 (ii) to see that ϕ(E1d)\phi(E_{1d}) is trivial, so also the conjugates ϕ(Eij)\phi(E_{ij}) are. As the EijE_{ij} generate SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) the result follows. For (d,m)=(3,1)(d,m)=(3,1) we use that SL3(𝐙)\mathrm{SL}_{3}(\mathbf{Z}) is perfect (see 2.1) and GL1(𝐂)\mathrm{GL}_{1}(\mathbf{C}) is abelian. For (d,m)=(3,2)(d,m)=(3,2) we apply the first part of 4.4: in case (i) we proceed as for d=4d=4 and the case (ii) is ruled out because the images of E12E_{12} and E13E_{13} are not conjugate. ∎

Corollary 4.6.

Fix d3d\geq 3 and a nontrivial homomorphism ϕ:SLd(𝐙)SLd(𝐂)\phi\colon\mathrm{SL}_{d}(\mathbf{Z})\rightarrow\mathrm{SL}_{d}(\mathbf{C}).

  1. (i)

    If d4d\geq 4, then for all iji\neq j the matrix ϕ(Eij)\phi(E_{ij}) is unipotent and ϕ(Eij)id\phi(E_{ij})-\mathrm{id} has rank 11. Moreover, after possibly precomposing ϕ\phi with ()t(-)^{-t}, the matrices ϕ(E1d),,ϕ(Ed1,d)\phi(E_{1d}),\ldots,\phi(E_{d-1,d}) all have the same fixed set.

  2. (ii)

    If d=3d=3, then the same conclusion holds under the additional assumption im(ϕ)SLd(𝐙)\mathrm{im}(\phi)\subset\mathrm{SL}_{d}(\mathbf{Z}).

Proof.

We begin with two observations based on the fact that EijSLd(𝐙)E_{ij}\in\mathrm{SL}_{d}(\mathbf{Z}) is for all iji\neq j conjugate to E1dE_{1d}. Firstly, to show the first claim of (i) and (ii), it suffices to consider ϕ(E1d)\phi(E_{1d}). Secondly, ϕ(E1d)\phi(E_{1d}) is nontrivial since otherwise ϕ\phi were trivial as SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) is generated by the EijE_{ij}.

In the case d4d\geq 4, it suffices to prove that ϕ(E1d)\phi(E_{1d}) is not a scalar, for then everything follows from 4.2, using that E1dE_{1d} is conjugate in SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) to EijE_{ij} for any iji\neq j. If ϕ(E1d)\phi(E_{1d}) were a scalar, then all ϕ(Eij)\phi(E_{ij}) are scalars, so ϕ\phi would have image in scalar matrices because the EijE_{ij} generate SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}). But since E1dE_{1d} is a commutator and scalar matrices commute, this would imply ϕ(E1d)=[ϕ(E12),ϕ(E2d)]=id\phi(E_{1d})=[\phi(E_{12}),\phi(E_{2d})]=\mathrm{id}, which is not the case.

Next we consider the case d=3d=3. To show the case m=3m=3, for which we imposed the additional assumption im(ϕ)SLd(𝐙)\mathrm{im}(\phi)\leq\mathrm{SL}_{d}(\mathbf{Z}). It suffices by 4.2 to prove that the nontrivial matrix ϕ(E13)\phi(E_{13}) is unipotent which we prove by contradiction. We consider the restriction of ϕ\phi to E12,E23𝐔3\langle E_{12},E_{23}\rangle\cong\mathbf{U}_{3} and consult the classification in 4.4. Since we assumed that ϕ(E13)id\phi(E_{13})\neq\mathrm{id} is not unipotent, we do not need to consider the case (i). Cases (ii) and (iii) of 4.4 can be excluded by showing that for these representations the matrices ϕ(E12),ϕ(E23),ϕ(E13)\phi(E_{12}),\phi(E_{23}),\phi(E_{13}) are not all conjugate in SL3(𝐙)\mathrm{SL}_{3}(\mathbf{Z}). In almost all cases this can be seen considering their eigenvalues, except in the case

ϕ(E12)=(100010001)ϕ(E23)=(010100001).\phi(E_{12})=\left(\begin{array}[]{ccc}1&0&0\\ 0&-1&0\\ 0&0&-1\end{array}\right)\>\>\>\>\>\>\phi(E_{23})=\left(\begin{array}[]{ccc}0&1&0\\ 1&0&0\\ 0&0&-1\end{array}\right).

Also these matrices are not conjugate in SL3(𝐙)\mathrm{SL}_{3}(\mathbf{Z}) which one can see by reducing modulo 22. ∎

Theorem 4.7.

Fix d3d\geq 3 and a nontrivial homomorphism ϕ:SLd(𝐙)SLd(𝐙)\phi:\mathrm{SL}_{d}(\mathbf{Z})\rightarrow\mathrm{SL}_{d}(\mathbf{Z}). There exist linearly independent vectors v1,,vd𝐙dv_{1},\ldots,v_{d}\in\mathbf{Z}^{d} so that, after possibly after precomposing ϕ\phi with ()t(-)^{-t}, the image of ϕ\phi preserves the lattice Λ=𝐙{v1,,vd}\Lambda=\mathbf{Z}\{v_{1},\ldots,v_{d}\} and for all ASLd(𝐙)A\in\mathrm{SL}_{d}(\mathbf{Z}) the matrix of the restriction ϕ(A)|Λ\phi(A)|_{\Lambda} with respect to the basis v1,,vdv_{1},\ldots,v_{d} is AA.

Remark 4.8.

One might suspect that given ϕ:SLd(𝐙)SLd(𝐂)\phi:\mathrm{SL}_{d}(\mathbf{Z})\rightarrow\mathrm{SL}_{d}(\mathbf{C}) there exists a basis v1,,vdv_{1},\ldots,v_{d} for 𝐂d\mathbf{C}^{d} so that the same conclusion of 4.7 holds (this is claimed in the MathOverflow post mentioned in Remark 4.3). This is not the case. For example, there is a nontrivial representation ϕ:SL3(𝐙)SL3(𝐂)\phi:\mathrm{SL}_{3}(\mathbf{Z})\rightarrow\mathrm{SL}_{3}(\mathbf{C}) with finite image, constructed by setting

ϕ(E12)=(100010001)ϕ(E23)=(010100001)ϕ(E32)=(100001010)ϕ(E21)=(121217412121+741+741740)\begin{array}[]{llll}\phi(E_{12})=&\left(\begin{array}[]{ccc}1&0&0\\ 0&-1&0\\ 0&0&-1\end{array}\right)&\phi(E_{23})=&\left(\begin{array}[]{ccc}0&1&0\\ 1&0&0\\ 0&0&-1\end{array}\right)\\ \phi(E_{32})=&\left(\begin{array}[]{ccc}-1&0&0\\ 0&0&1\\ 0&1&0\end{array}\right)&\phi(E_{21})=&\left(\begin{array}[]{ccc}-\frac{1}{2}&-\frac{1}{2}&\frac{-1-\sqrt{-7}}{4}\\ -\frac{1}{2}&-\frac{1}{2}&\frac{1+\sqrt{-7}}{4}\\ \frac{-1+\sqrt{-7}}{4}&\frac{1-\sqrt{-7}}{4}&0\end{array}\right)\end{array}

and then defining ϕ(E13)[ϕ(E12),ϕ(E23)]\phi(E_{13})\coloneqq[\phi(E_{12}),\phi(E_{23})] and ϕ(E31)[ϕ(E32),ϕ(E21)]\phi(E_{31})\coloneqq[\phi(E_{32}),\phi(E_{21})]. One can then check directly that this extends to a morphism SL3(𝐙)SL3(𝐐(7))SL3(𝐂)\mathrm{SL}_{3}(\mathbf{Z})\rightarrow\mathrm{SL}_{3}(\mathbf{Q}(\sqrt{-7}))\subset\mathrm{SL}_{3}(\mathbf{C}) by checking that these matrices satisfy the relations in the standard presentation of SL3(𝐙)\mathrm{SL}_{3}(\mathbf{Z}) in terms of EijE_{ij} (see [Mil71, Corollary 10.3]). This peculiar representation has finite image because for each iji\neq j, the matrix ϕ(Eij)\phi(E_{ij}) has order 2, and the subgroup generated by {Eij2}\{E_{ij}^{2}\} has finite index in SL3(𝐙)\mathrm{SL}_{3}(\mathbf{Z}) by a general theorem of Tits [Tit76] (see also [Mei17, Theorem 3]).

Proof of Theorem 4.7.

Fix a nontrivial homomorphism ϕ:SLd(𝐙)SLd(𝐙)\phi:\mathrm{SL}_{d}(\mathbf{Z})\rightarrow\mathrm{SL}_{d}(\mathbf{Z}). We write uij=ϕ(Eij)u_{ij}=\phi(E_{ij}), considered as a matrix in SLd(𝐂)\mathrm{SL}_{d}(\mathbf{C}). After possibly precomposing ϕ\phi with ()t(-)^{-t}, we know from 4.6, that for iji\neq j, the matrix uiju_{ij} is unipotent and uijidu_{ij}-\mathrm{id} has rank 11, and that H1d=H2d==Hd1,dH_{1d}=H_{2d}=\cdots=H_{d-1,d} where Hij𝐂dH_{ij}\leq\mathbf{C}^{d} be the fixed set of the matrix uiju_{ij} for iji\neq j. Note that each HijH_{ij} is (d1)(d-1)-dimensional, since uijidu_{ij}-\mathrm{id} has rank 1. Using the fact that for each fixed 1kd1\leq k\leq d, the matrices E1k,E2k,,EdkE_{1k},E_{2k},\ldots,E_{dk} (skipping EkkE_{kk}) are simultaneously conjugate to E1d,,Ed1,dE_{1d},\ldots,E_{d-1,d}, we find that also the d1d-1 hyperplanes H1k,H2k,,HdkH_{1k},H_{2k},\ldots,H_{dk} (skipping HkkH_{kk}) all agree. We abbreviate this hyperplane by HkH_{k}. Next we claim that the intersection of hyperplanes Li=H1H^iHdL_{i}=H_{1}\cap\cdots\cap\widehat{H}_{i}\cap\cdots\cap H_{d} for 1id1\leq i\leq d are all lines. For this it suffices to show that H1HdH_{1}\cap\cdots\cap H_{d} is trivial. Assume by contradiction that this intersection is nontrivial. By construction, it is the common fixed set for the uiju_{ij} for all iji\neq j, so it is in fact fixed by the whole image of ϕ\phi since the uij=ϕ(Eij)u_{ij}=\phi(E_{ij}) generate the image because the EijE_{ij} generate SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}). Moreover, since the HiH_{i} are defined over 𝐐\mathbf{Q}, also L:-H1Hd𝐙dL\coloneq H_{1}\cap\cdots\cap H_{d}\cap\mathbf{Z}^{d} is nontrivial, so the free abelian group 𝐙d/L\mathbf{Z}^{d}/L has rank <d<d. Combining this with 4.5, we see that the morphism SLd(𝐙)SL(𝐙d/L)\mathrm{SL}_{d}(\mathbf{Z})\rightarrow\mathrm{SL}(\mathbf{Z}^{d}/L) induced by ϕ\phi is trivial, so ϕ\phi factors over the additive group Hom(𝐙d/L,L)\mathrm{Hom}(\mathbf{Z}^{d}/L,L). The latter is abelian, so ϕ\phi must be trivial since SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) is perfect (see 2.1). This contradicts our choice of ϕ\phi.

Claim. The image of uijidu_{ij}-\mathrm{id} is LiL_{i}.

Proof of Claim. For definiteness, we prove the statement for u1du_{1d}. Since u1didu_{1d}-\mathrm{id} has rank 11 and L1=H2HdL_{1}=H_{2}\cap\ldots\cap H_{d} is 11-dimensional, it suffices to show that the image of u1didu_{1d}-\mathrm{id} is contained in HjH_{j} for all j1j\neq 1. Recall that Hj=H1jH_{j}=H_{1j} is the fixed set of u1ju_{1j}. Since u1ju_{1j} commutes with u1du_{1d}, the matrix u1ju_{1j} preserves the image of u1didu_{1d}-\mathrm{id}, but since this image is only one dimensional, it is an eigenspace for u1ju_{1j}, which implies im(u1did)H1j\mathrm{im}(u_{1d}-\mathrm{id})\subset H_{1j} since u1ju_{1j} is unipotent. This proves the claim.

Now we construct the basis v1,,vdv_{1},\ldots,v_{d}. Fix a nonzero vector vdLdv_{d}\in L_{d} which we may choose to be an integer vector as LdL_{d} is defined over 𝐐\mathbf{Q} since ϕ\phi has image in SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}). Now define inductively vi=(ui,i+1id)(vi+1)Liv_{i}=(u_{i,i+1}-\mathrm{id})(v_{i+1})\in L_{i}. Note that the viv_{i} are integer vectors as ui,i+1=ϕ(Ei,i+1)SLd(𝐙)u_{i,i+1}=\phi(E_{i,i+1})\in\mathrm{SL}_{d}(\mathbf{Z}). Moreover, each viv_{i} is nonzero: if viv_{i} were trivial then vi+1v_{i+1} would be contained in Li+1Hi+1=H1HdL_{i+1}\cap H_{i+1}=H_{1}\cap\cdots\cap H_{d} which we saw above is trivial, so we get vi+1=0v_{i+1}=0 and inductively vd=0v_{d}=0 which is not true. Now we examine what properties the vectors v1,,vdv_{1},\ldots,v_{d} have. First observe that they form a basis for 𝐂d\mathbf{C}^{d}, by the general fact that if H1,,HdH_{1},\ldots,H_{d} are hyperplanes of 𝐂d\mathbf{C}^{d} with trivial intersection, then a choice of nonzero vector from each of the lines Li=H1Hi^HdL_{i}=H_{1}\cap\cdots\cap\widehat{H_{i}}\cap\cdots\cap H_{d} gives a basis for 𝐂d\mathbf{C}^{d}. By construction, with respect to the basis v1,,vdv_{1},\ldots,v_{d}, the matrix of ui,i+1=ϕ(Ei,i+1)u_{i,i+1}=\phi(E_{i,i+1}) is Ei,i+1E_{i,i+1}. Now using the commutator relations in 𝐔d\mathbf{U}_{d}, we conclude that after this change of basis the restriction of ϕ\phi to upper triangular matrices is the inclusion, so to finish the proof suffices to show the same for the lower triangular matrices since SLd(𝐙)\mathrm{SL}_{d}(\mathbf{Z}) is generated by upper and lower triangular matrices. As for upper triangular matrices, it suffices to consider Ei+1,iE_{i+1,i} for every ii. By construction, ϕ(Ei+1,i)\phi(E_{i+1,i}) has fixed set v1,,vi^,,vd\langle v_{1},\ldots,\widehat{v_{i}},\ldots,v_{d}\rangle and (ϕ(Ei+1,i)id)(vi)=aivi+1(\phi(E_{i+1,i})-\mathrm{id})(v_{i})=a_{i}\cdot v_{i+1} for some scalar aia_{i}, so we are left to show ai=1a_{i}=1. This follows from the braid relation Ei,i+11Ei+1,iEi,i+11=Ei+1,iEi,i+11Ei+1,iE_{i,i+1}^{-1}\>E_{i+1,i}\>E_{i,i+1}^{-1}=E_{i+1,i}\>E_{i,i+1}^{-1}\>E_{i+1,i}. ∎

Proof of Theorem E.

Fix a nontrivial homomorphism ϕ:SLd(𝐙)SLd(𝐙)\phi\colon\mathrm{SL}_{d}(\mathbf{Z})\rightarrow\mathrm{SL}_{d}(\mathbf{Z}) and let v1,,vd𝐙dv_{1},\ldots,v_{d}\in\mathbf{Z}^{d} be the linearly independent vectors promised by 4.7, so that possibly after precomposing ϕ\phi with ()t(-)^{-t}, the matrix ϕ(A)\phi(A) for ASLd(𝐙)A\in\mathrm{SL}_{d}(\mathbf{Z}) preserves the lattice Λ=𝐙{v1,,vd}\Lambda=\mathbf{Z}\{v_{1},\ldots,v_{d}\}, and the restriction ϕ(A)|Λ\phi(A)|_{\Lambda} is represented by the matrix AA when written in the basis v1,,vdv_{1},\ldots,v_{d}. In particular, this has as consequence that every orientation-preserving automorphism of Λ𝐙d\Lambda\leq\mathbf{Z}^{d} extends to an orientation-preserving automorphism of 𝐙d\mathbf{Z}^{d}. We claim that this in turn implies Λ=𝐙d\Lambda=\ell\mathbf{Z}^{d} for some >0\ell>0. Dividing the basis by \ell, this would show that we can choose v1,,vdv_{1},\ldots,v_{d} to form a basis of 𝐙d\mathbf{Z}^{d}, so ϕ\phi is given by conjugation by an element of GLd(𝐙)\mathrm{GL}_{d}(\mathbf{Z}). That Λ=𝐙d\Lambda=\ell\cdot\mathbf{Z}^{d} for some >0\ell>0 follows from two facts: (a) for every non-characteristic subgroup L𝐙dL\subset\mathbf{Z}^{d} of full rank, there exists an (orientation-preserving) automorphism of LL that does not extend to 𝐙d\mathbf{Z}^{d}, so Λ\Lambda has to be characteristic, and (b) every characteristic subgroup L𝐙dL\leq\mathbf{Z}^{d} of full rank has the form 𝐙d\ell\cdot\mathbf{Z}^{d} for some >0\ell>0. To see these two facts, we fix a subgroup L𝐙dL\leq\mathbf{Z}^{d} of full rank. By the elementary divisor theorem, there is a basis b1,,bdb_{1},\ldots,b_{d} of 𝐙d\mathbf{Z}^{d} and natural numbers 1,,d\ell_{1},\ldots,\ell_{d} such that 1b1,,dbd\ell_{1}\cdot b_{1},\ldots,\ell_{d}\cdot b_{d} is a basis of LL. If LL is non-characteristic, then ij\ell_{i}\neq\ell_{j} for some ii and jj (since 𝐙d𝐙d\ell\cdot\mathbf{Z}^{d}\leq\mathbf{Z}^{d} is clearly characteristic), so the automorphism of LL that interchanges ibi\ell_{i}\cdot b_{i} and jbj\ell_{j}\cdot b_{j} does not extend to 𝐙d\mathbf{Z}^{d} (by interchanging a second pair of basis vectors we also find an orientation-preserving example of such an automorphism). This shows (a). Moreover, if we assume ij\ell_{i}\neq\ell_{j} for some ii and jj, then the automorphism of 𝐙d\mathbf{Z}^{d} that interchanges bib_{i} and bjb_{j} does not restrict to LL, so LL cannot be characteristic. This shows (b). ∎

References

  • [Arn14] V. I. Arnold, On some topological invariants of algebraic functions, Vladimir I. Arnold - Collected Works: Hydrodynamics, Bifurcation Theory, and Algebraic Geometry 1965-1972 (A. B. Givental, B. A. Khesin, A. N. Varchenko, V. A. Vassiliev, and O. Y. Viro, eds.), Springer Berlin Heidelberg, Berlin, Heidelberg, 2014, pp. 199–221.
  • [Beh20] M. Behrens, Topological modular and automorphic forms, Handbook of homotopy theory, CRC Press/Chapman Hall Handb. Math. Ser., CRC Press, Boca Raton, FL, 2020, pp. 221–261.
  • [BFH20] A. Brown, D. Fisher, and S. Hurtado, Zimmer’s conjecture for actions of SL(m,){\rm SL}(m,\mathbb{Z}), Invent. Math. 221 (2020), no. 3, 1001–1060.
  • [Bre67] G. E. Bredon, A Π\Pi_{\ast}-module structure for Θ\Theta_{\ast} and applications to transformation groups, Ann. of Math. (2) 86 (1967), 434–448.
  • [Bro94] K. S. Brown, Cohomology of groups, Graduate Texts in Mathematics, vol. 87, Springer-Verlag, New York, 1994, Corrected reprint of the 1982 original.
  • [Bru68] G. Brumfiel, On the homotopy groups of BPL{\rm BPL} and PL/O{\rm PL/O}, Ann. of Math. (2) 88 (1968), 291–311.
  • [Bru69] by same author, On the homotopy groups of BPLBPL and PL/OPL/O. II, Topology 8 (1969), 305–311.
  • [Bru70] by same author, The homotopy groups of BPLBPL and PL/OPL/O. III, Michigan Math. J. 17 (1970), 217–224.
  • [BT21] M. Bustamante and B. Tshishiku, Symmetries of exotic aspherical space forms, 2021, arXiv:2109.09196.
  • [CT22] L. Chen and B. Tshishiku, Nielsen realization for sphere twists on 3-manifolds, 2022, to appear in Israel J. Math.
  • [FM22] D. Fisher and K. Melnick, Smooth and analytic actions of SL(n,𝐑)SL(n,{\bf R}) and SL(n,𝐙)SL(n,{\bf Z}) on closed nn-dimensional manifolds, 2022, to appear in Kyoto J. Math.
  • [FQ90] M. H. Freedman and F. Quinn, Topology of 4-manifolds, Princeton Mathematical Series, vol. 39, Princeton University Press, Princeton, NJ, 1990.
  • [Fra73] D. Frank, The signature defect and the homotopy of BPL{\rm BPL} and PL/O{\rm PL/O}, Comment. Math. Helv. 48 (1973), 525–530.
  • [Hat76] A. Hatcher, Homeomorphisms of sufficiently large P2P^{2}-irreducible 33-manifolds, Topology 15 (1976), no. 4, 343–347.
  • [Hat78] A. E. Hatcher, Concordance spaces, higher simple-homotopy theory, and applications, Algebraic and geometric topology (Proc. Sympos. Pure Math., Stanford Univ., Stanford, Calif., 1976), Part 1, Proc. Sympos. Pure Math., XXXII, Amer. Math. Soc., Providence, R.I., 1978, pp. 3–21.
  • [Hat83] by same author, A proof of the Smale conjecture, Diff(S3)O(4){\rm Diff}(S^{3})\simeq{\rm O}(4), Ann. of Math. (2) 117 (1983), no. 3, 553–607.
  • [HHR16] M. A. Hill, M. J. Hopkins, and D. C. Ravenel, On the nonexistence of elements of Kervaire invariant one, Ann. of Math. (2) 184 (2016), no. 1, 1–262.
  • [HLLRW21] F. Hebestreit, M. Land, W. Lück, and O. Randal-Williams, A vanishing theorem for tautological classes of aspherical manifolds, Geom. Topol. 25 (2021), no. 1, 47–110.
  • [HS76] W. C. Hsiang and R. W. Sharpe, Parametrized surgery and isotopy, Pacific J. Math. 67 (1976), no. 2, 401–459.
  • [HW69] W.-C. Hsiang and C. T. C. Wall, On homotopy tori. II, Bull. London Math. Soc. 1 (1969), 341–342.
  • [IWX20] D. C. Isaksen, G. Wang, and Z. Xu, Stable homotopy groups of spheres, Proc. Natl. Acad. Sci. USA 117 (2020), no. 40, 24757–24763.
  • [Kir97] R. Kirby, Problems in low-dimensional topology, Geometric topology (Athens, GA, 1993) (Rob Kirby, ed.), AMS/IP Stud. Adv. Math., vol. 2, Amer. Math. Soc., Providence, RI, 1997, pp. 35–473.
  • [KM63] M. A. Kervaire and J. W. Milnor, Groups of homotopy spheres. I, Ann. of Math. (2) 77 (1963), 504–537.
  • [Kor02] M. Korkmaz, Low-dimensional homology groups of mapping class groups: a survey, Turkish J. Math. 26 (2002), no. 1, 101–114.
  • [Kra21] M. Krannich, On characteristic classes of exotic manifold bundles, Math. Ann. 379 (2021), no. 1-2, 1–21.
  • [Kre79] M. Kreck, Isotopy classes of diffeomorphisms of (k1)(k-1)-connected almost-parallelizable 2k2k-manifolds, Algebraic topology, Aarhus 1978 (Proc. Sympos., Univ. Aarhus, Aarhus, 1978), Lecture Notes in Math., vol. 763, Springer, Berlin, 1979, pp. 643–663.
  • [KS77] R. C. Kirby and L. C. Siebenmann, Foundational essays on topological manifolds, smoothings, and triangulations, Princeton University Press, Princeton, N.J.; University of Tokyo Press, Tokyo, 1977, With notes by John Milnor and Michael Atiyah, Annals of Mathematics Studies, No. 88.
  • [Law76] T. Lawson, Homeomorphisms of Bk×TnB^{k}\times T^{n}, Proc. Amer. Math. Soc. 56 (1976), 349–350.
  • [Law73] T. C. Lawson, Remarks on the pairings of Bredon, Milnor, and Milnor-Munkres-Novikov, Indiana Univ. Math. J. 22 (1972/73), 833–843.
  • [Lev70] J. Levine, Inertia groups of manifolds and diffeomorphisms of spheres, Amer. J. Math. 92 (1970), 243–258.
  • [Lev85] J. P. Levine, Lectures on groups of homotopy spheres, Algebraic and geometric topology (New Brunswick, N.J., 1983), Lecture Notes in Math., vol. 1126, Springer, Berlin, 1985, pp. 62–95.
  • [LR81] K. B. Lee and F. Raymond, Topological, affine and isometric actions on flat Riemannian manifolds, J. Differential Geometry 16 (1981), no. 2, 255–269.
  • [Mat17] MathOverflow, SLn()SL_{n}(\mathbb{Z}) is co-hopfian, 2017, https://mathoverflow.net/q/272618 (version: 2017-06-20).
  • [Mei17] C. Meiri, Generating pairs for finite index subgroups of SL(n,){\rm SL}(n,\mathbb{Z}), J. Algebra 470 (2017), 420–424.
  • [Men65] J. L. Mennicke, Finite factor groups of the unimodular group, Ann. of Math. (2) 81 (1965), 31–37.
  • [Mil71] J. Milnor, Introduction to algebraic KK-theory, Annals of Mathematics Studies, No. 72, Princeton University Press, Princeton, N.J.; University of Tokyo Press, Tokyo, 1971.
  • [O’M66] O. T. O’Meara, The automorphisms of the linear groups over any integral domain, J. Reine Angew. Math. 223 (1966), 56–100.
  • [Ser03] J.-P. Serre, Trees, Springer Monographs in Mathematics, Springer-Verlag, Berlin, 2003, Translated from the French original by John Stillwell, Corrected 2nd printing of the 1980 English translation.
  • [Tit76] J. Tits, Systèmes générateurs de groupes de congruence, C. R. Acad. Sci. Paris Sér. A-B 283 (1976), no. 9, Ai, A693–A695.
  • [vdK75] W. van der Kallen, The Schur multipliers of SL(3,𝐙){\rm SL}(3,{\bf Z}) and SL(4,𝐙){\rm SL}(4,{\bf Z}), Math. Ann. 212 (1974/75), 47–49.
  • [Wal68] F. Waldhausen, On irreducible 33-manifolds which are sufficiently large, Ann. of Math. (2) 87 (1968), 56–88.
  • [Wei97] S. Weinberger, SL(n,){\rm SL}(n,\mathbb{Z}) cannot act on small tori, Geometric topology (Athens, GA, 1993), AMS/IP Stud. Adv. Math., vol. 2, Amer. Math. Soc., Providence, RI, 1997, pp. 406–408.