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

Functoriality for Higher Rho Invariants of Elliptic Operators

Hao Guo Department of Mathematics, Texas A&M University [email protected] Zhizhang Xie Department of Mathematics, Texas A&M University [email protected]  and  Guoliang Yu Department of Mathematics, Texas A&M University [email protected]
Abstract.

Let NN be a closed spin manifold with positive scalar curvature and DND_{N} the Dirac operator on NN. Let M1M_{1} and M2M_{2} be two Galois covers of NN such that M2M_{2} is a quotient of M1M_{1}. Then the quotient map from M1M_{1} to M2M_{2} naturally induces maps between the geometric CC^{*}-algebras associated to the two manifolds. We prove, by a finite-propagation argument, that the maximal higher rho invariants of the lifts of DND_{N} to M1M_{1} and M2M_{2} behave functorially with respect to the above quotient map. This can be applied to the computation of higher rho invariants, along with other related invariants.

2010 Mathematics Subject Classification:
46L80, 58B34, 53C20
H.G. is partially supported by NSF 1564398. Z.X. is partially supported by NSF 1500823 and 1800737. G.Y. is partially supported by NSF 1700021, 1564398, and the Simons Fellows Program

1. Introduction

An elliptic differential operator on a closed manifold has a Fredholm index. When such an operator is lifted to a covering space, one can, by taking into account the group of symmetries, define a far-reaching generalization of the Fredholm index, called the higher index [1, 2, 6, 13, 26]. The higher index serves as an obstruction to the existence of invariant metrics of positive scalar curvature. In the case that such a metric exists, so that the higher index of the lifted operator vanishes, a secondary invariant called the higher rho invariant [19, 12] can be defined. The higher rho invariant is an obstruction to the inverse of the operator being local [4]. For some recent applications of the higher index and higher rho invariant to problems in geometry and topology, we refer the reader to [5, 22, 23, 25, 27, 28, 29].

The main purpose of this paper is to prove that the higher rho invariant behaves functorially under appropriate maps between covering spaces.

More precisely, suppose M1M_{1} and M2M_{2} are two Galois covers of a closed spin manifold NN with deck transformation groups Γ1\Gamma_{1} and Γ2Γ1/H\Gamma_{2}\cong\Gamma_{1}/H, where HH is a normal subgroup of Γ1\Gamma_{1}. The natural projection π:M1M2\pi\colon M_{1}\rightarrow M_{2} induces a family of maps between various geometric CC^{*}-algebras associated to M1M_{1} and M2M_{2}, called folding maps, which will be reviewed in section 2. In particular, there is a folding map

Ψ:Cmax(M1)Γ1\displaystyle\Psi\colon C^{*}_{\textnormal{max}}(M_{1})^{\Gamma_{1}} Cmax(M2)Γ2\displaystyle\rightarrow C^{*}_{\textnormal{max}}(M_{2})^{\Gamma_{2}}

between maximal equivariant Roe algebras on M1M_{1} and M2M_{2} with the property that it preserves small propagation of operators. Consequently, Ψ\Psi induces a map

ΨL,0:CL,0,max(M1)Γ1CL,0,max(M2)Γ2\Psi_{L,0}\colon C^{*}_{L,0,\textnormal{max}}(M_{1})^{\Gamma_{1}}\rightarrow C^{*}_{L,0,\textnormal{max}}(M_{2})^{\Gamma_{2}}

at the level of obstruction algebras. The main result of this paper is that the KK-theoretic map induced by ΨL,0\Psi_{L,0} relates the maximal higher rho invariants of Dirac operators on M1M_{1} and M2M_{2}:

Theorem 1.1.

Let NN be a closed, spin Riemannian manifold with positive scalar curvature. Let DND_{N} be the Dirac operator on NN. Let M1M_{1} and M2M_{2} be Galois covers of MM with deck transformation groups Γ1\Gamma_{1} and Γ2Γ1/H\Gamma_{2}\cong\Gamma_{1}/H respectively, for a normal subgroup HH of Γ1\Gamma_{1}. Let D1D_{1} and D2D_{2} be the lifts of the Dirac operator on NN to M1M_{1} and M2M_{2} respectively. Then

(ΨL,0)(ρmax(D1))=ρmax(D2)K(CL,0,max(M2)Γ2),(\Psi_{L,0})_{*}(\rho_{\textnormal{max}}(D_{1}))=\rho_{\textnormal{max}}(D_{2})\in K_{\bullet}(C^{*}_{L,0,\textnormal{max}}(M_{2})^{\Gamma_{2}}),

where (ΨL,0)(\Psi_{L,0})_{*} is the map on KK-theory induced by ΨL,0\Psi_{L,0} and ρmax(Dj)\rho_{\textnormal{max}}(D_{j}) is the maximal higher rho invariant of DjD_{j}.

We also generalize this result to the non-cocompact setting using the same method of proof – see section 7 of this paper.

Theorem 1.1 is useful for computing higher rho and related invariants. It has been applied in the recent work of Wang, Xie, and Yu [24] to compute the delocalized eta invariant on a covering space of a closed manifold by showing that, under suitable geometric conditions, it can be approximated by delocalized eta invariants on finite-sheeted covers, which are more computable. We mention two of their results.

Let MM be a closed spin manifold equipped with a positive scalar curvature metric. First, the authors proved that given a finitely generated discrete group Γ\Gamma and a sequence {Γi}i\{\Gamma_{i}\}_{i\in\mathbb{N}} of finite-index normal subgroups of Γ\Gamma that distinguishes, in a suitable sense ([25, Definition 2.3]), a given non-trivial conjugacy class in Γ\Gamma, the sequence of associated delocalized eta invariants on the Γi\Gamma_{i}-Galois covers of MM stabilizes, under the assumption that the maximal Baum-Connes assembly map for Γ\Gamma is a rational isomorphism. We refer to [25, Theorem 1.1] for more details.

Second, the authors proved that if D~\widetilde{D} is the Dirac operator associated to the Γ\Gamma-Galois cover of MM, then if the spectral gap of D~\widetilde{D} at zero is sufficiently large, the delocalized eta invariant of D~\widetilde{D} is equal to the limit of those of Dirac operators associated to the Γi\Gamma_{i}-Galois covers of MM [25, Theorem 1.4]. In particular, this is true if the group Γ\Gamma has subexponential growth [25, Corollary 1.5].

Finally, we mention that in a new preprint [16] by Liu, Tang, Xie, Yao, and Yu, the authors apply Theorem 1.1 to study rigidity of relative eta invariants. These invariants are obtained by pairing the higher rho invariant with certain traces, similar to the way in which the delocalized eta invariant is related to the higher rho invariant through a trace [28].

Overview.

The paper is organized as follows. We begin in section 2 by recalling the geometric and operator-algebraic setup we work with. In section 3 we define the folding maps and establish some of their basic properties. In section 4 we develop the analytical properties of the wave operator in the maximal setting. These tools are then put to use in section 5, where we give a new proof of the functoriality property of the maximal equivariant higher index. This serves as an intermediate step towards Theorem 1.1, which is proved in section 6. In section 7 we provide generalizations of our results to the non-cocompact setting.

Acknowledgements

The authors are grateful to Peter Hochs for pointing us to the reference [17, Theorem 1.1] of Alain Valette for functoriality of the maximal higher index.


2. Preliminaries

In this section, we fix some notation before introducing the necessary operator-algebraic background and geometric setup for our results.

2.1. Notation

For XX a Riemannian manifold, we write B(X)B(X), Cb(X)C_{b}(X), C0(X)C_{0}(X), and Cc(X)C_{c}(X) to denote the CC^{*}-algebras of complex-valued functions on XX that are, respectively: bounded Borel, bounded continuous, continuous and vanishing at infinity, and continuous with compact support. A superscript ‘’ may be added where appropriate to indicate the additional requirement of smoothness.

We write dXd_{X} for the Riemannian distance function on XX and 𝟙S\mathbbm{1}_{S} for the characteristic function of a subset SXS\subseteq X.

For any CC^{*}-algebra AA, we denote its unitization by A+A^{+}, its multiplier algebra by (A)\mathcal{M}(A), and view AA as an ideal of (A)\mathcal{M}(A).

The action of a group GG on XX naturally induces a GG-action on spaces of functions on XX as follows: given a function ff on XX and gGg\in G, define gfg\cdot f by gf(x)=f(g1x)g\cdot f(x)=f(g^{-1}x). More generally, for a section ss of a Γ\Gamma-vector bundle over XX, the section gsg\cdot s is defined by gs(x)=g(s(g1x))g\cdot s(x)=g(s(g^{-1}x)). We say that an operator on sections of a bundle is GG-equivariant if it commutes with the GG-action.

2.2. Geometric CC^{*}-algebras

We now recall the notions of geometric modules and their associated CC^{*}-algebras. Throughout this subsection, XX is a Riemannian manifold equipped with a proper isometric action by a discrete group GG.

Definition 2.1.

An XX-GG-module is a separable Hilbert space \mathcal{H} equipped with a non-degenerate *-representation ρ:C0(X)()\rho\colon C_{0}(X)\rightarrow\mathcal{B}(\mathcal{H}) and a unitary representation U:G𝒰()U\colon G\rightarrow\mathcal{U}(\mathcal{H}) such that for all fC0(X)f\in C_{0}(X) and gGg\in G, we have Ugρ(f)Ug=ρ(gf)U_{g}\rho(f)U_{g}^{*}=\rho(g\cdot f).

For brevity, we will omit ρ\rho from the notation when it is clear from context.

Definition 2.2.

Let \mathcal{H} be an XX-GG-module and T()T\in\mathcal{B}(\mathcal{H}).

  • The support of TT, denoted supp(T)\textnormal{supp}(T), is the complement of all (x,y)X×X(x,y)\in X\times X for which there exist f1,f2C0(X)f_{1},f_{2}\in C_{0}(X) such that f1(x)0f_{1}(x)\neq 0, f2(y)0f_{2}(y)\neq 0, and

    f1Tf2=0;f_{1}Tf_{2}=0;
  • The propagation of TT is the extended real number

    prop(T)=sup{dX(x,y)|(x,y)supp(T)};\textnormal{prop}(T)=\sup\{d_{X}(x,y)\,|\,(x,y)\in\textnormal{supp}(T)\};
  • TT is locally compact if fTfT and Tf𝒦()Tf\in\mathcal{K}(\mathcal{H}) for all fC0(X)f\in C_{0}(X);

  • TT is GG-equivariant if UgTUg=TU_{g}TU_{g}^{*}=T for all gGg\in G;

The equivariant algebraic Roe algebra for \mathcal{H}, denoted [X;]G\mathbb{C}[X;\mathcal{H}]^{G}, is the *-subalgebra of ()\mathcal{B}(\mathcal{H}) consisting of GG-equivariant, locally compact operators with finite propagation.

We will work with the maximal completion of the equivariant algebraic Roe algebra. To ensure that this completion is well-defined, we require that the module \mathcal{H} satisfy an additional admissibility condition. To define what this means, we need the following fact: if HH is a Hilbert space and ρ:C0(X)(H)\rho\colon C_{0}(X)\rightarrow\mathcal{B}(H) is a non-degenerate *-representation, then ρ\rho extends uniquely to a *-representation ρ~:B(X)(H)\widetilde{\rho}\colon B(X)\rightarrow\mathcal{B}(H) subject to the property that, for a uniformly bounded sequence in B(X)B(X) converging pointwise, the corresponding sequence in (H)\mathcal{B}(H) converges in the strong topology.

Definition 2.3 ([30]).

Let \mathcal{H} be an XX-GG-module as in Definition 2.1. We say that \mathcal{H} is admissible if:

  1. (i)

    For any non-zero fC0(X)f\in C_{0}(X) we have ρ(f)𝒦()\rho(f)\notin\mathcal{K}(\mathcal{H});

  2. (ii)

    For any finite subgroup FF of GG and any FF-invariant Borel subset EXE\subseteq X, there is a Hilbert space HH^{\prime} equipped with the trivial FF-representation such that ρ~(𝟙E)Hl2(F)H\widetilde{\rho}(\mathbbm{1}_{E})H^{\prime}\cong l^{2}(F)\otimes H^{\prime} as FF-representations, where ρ~\widetilde{\rho} is defined by extending ρ\rho as above.

If an XX-GG-module \mathcal{H} is admissible, we will write [X]G\mathbb{C}[X]^{G} in place of [X;]G\mathbb{C}[X;\mathcal{H}]^{G}, for the reason that [X;]G\mathbb{C}[X;\mathcal{H}]^{G} is independent of the choice of admissible module – see [26, Chapter 5].

Remark 2.4.

When GG acts freely and properly on XX, the Hilbert space L2(X)L^{2}(X) is an admissible XX-GG-module. In the case that the action is not free, L2(X)L^{2}(X) can always be embedded into a larger admissible module.

Definition 2.5.

The maximal norm of an operator T[X]GT\in\mathbb{C}[X]^{G} is

||T||maxsupϕ,H{ϕ(T)(H)|ϕ:[X]G(H) is a -representation}.||T||_{\textnormal{max}}\coloneqq\sup_{\phi,H^{\prime}}\left\{\|\phi(T)\|_{\mathcal{B}(H^{\prime})}\,|\,\phi\colon\mathbb{C}[X]^{G}\rightarrow\mathcal{B}(H^{\prime})\textnormal{ is a $*$-representation}\right\}.

The maximal equivariant Roe algebra of MjM_{j}, denoted Cmax(X)GC^{*}_{\text{max}}(X)^{G}, is the completion of [X]G\mathbb{C}[X]^{G} in the norm ||||max||\cdot||_{\textnormal{max}}.

Remark 2.6.

To make sense of Definition 2.5 for general XX and GG, one first needs to establish finiteness of the quantity max\|\cdot\|_{\textnormal{max}}. It was shown in [8] that if GG acts on XX freely and properly with compact quotient, the norm max\|\cdot\|_{\textnormal{max}} is finite. This was generalized in [11] to the case when XX has bounded geometry and the GG-action satisfies a suitable geometric assumption.

Remark 2.7.

Equivalently, one can obtain Cmax(X)GC^{*}_{\text{max}}(X)^{G} by taking the analogous maximal completion of the subalgebra 𝒮G\mathcal{S}^{G} of [X]G\mathbb{C}[X]^{G} consisting of those operators given by smooth Schwartz kernels.

Definition 2.8.

Consider the *-algebra LL of functions f:[0,)Cmax(X)Gf\colon[0,\infty)\rightarrow C^{*}_{\textnormal{max}}(X)^{G} that are uniformly bounded, uniformly continuous, and such that

prop(f(t))0as t.\textnormal{prop}(f(t))\rightarrow 0\quad\textnormal{as }t\rightarrow\infty.
  1. (i)

    The maximal equivariant localization algebra, denoted by CL,max(X)GC^{*}_{L,\textnormal{max}}(X)^{G}, is the CC^{*}-algebra obtained by completing LL with respect to the norm

    fsuptf(t)max;\|f\|\coloneqq\sup_{t}\|f(t)\|_{\textnormal{max}};
  2. (ii)

    The map L[X]GL\rightarrow\mathbb{C}[X]^{G} given by ff(0)f\mapsto f(0) extends to the evaluation map

    ev:CL,max(X)GCmax(X)G;\textnormal{ev}\colon C^{*}_{L,\textnormal{max}}(X)^{G}\rightarrow C^{*}_{\textnormal{max}}(X)^{G};
  3. (iii)

    The maximal equivariant obstruction algebra is CL,0,max(X)Gker(ev)C^{*}_{L,0,\textnormal{max}}(X)^{G}\coloneqq\ker(\textnormal{ev}).

2.3. Geometric setup

We will work with the following geometric setup.

Let (N,gN)(N,g_{N}) be a closed Riemannian manifold. Let DND_{N} be a first-order essentially self-adjoint elliptic differential operator on a bundle ENNE_{N}\rightarrow N. We will assume throughout that if NN is odd-dimensional then DND_{N} is an ungraded operator, while if NN is even-dimensional then DND_{N} is odd-graded with respect to a 2\mathbb{Z}_{2}-grading on ENE_{N}.

Let p1:M1Np_{1}\colon M_{1}\rightarrow N and p2:M2Np_{2}\colon M_{2}\rightarrow N be two Galois covers of NN with deck transformation groups Γ1\Gamma_{1} and Γ2\Gamma_{2} respectively. We will assume throughout this paper that Γ2Γ1/H\Gamma_{2}\cong\Gamma_{1}/H for some normal subgroup HH of Γ1\Gamma_{1}, so that M2M1/HM_{2}\cong M_{1}/H. Let π:M1M2\pi\colon M_{1}\rightarrow M_{2} be the projection map. Note that for j=1,2j=1,2, the group Γj\Gamma_{j} acts freely and properly on MjM_{j}.

For j=1j=1 or 22, let gjg_{j} be the lift of the Riemannian metric gNg_{N} to MjM_{j}. Let EjE_{j} be the pullback of EE along the covering map MjMM_{j}\rightarrow M, equipped with the natural Γj\Gamma_{j}-action. Since DD acts locally, it lifts to a Γj\Gamma_{j}-equivariant operator DjD_{j} on C(Ej)C^{\infty}(E_{j}).

We will apply the notions in subsection 2.2 with X=MjX=M_{j} and G=ΓjG=\Gamma_{j}. The Hilbert space L2(Ej)L^{2}(E_{j}), equipped with the natural Γj\Gamma_{j}-action on sections and the C0(Mj)C_{0}(M_{j})-representation defined by pointwise multiplication, is an MjM_{j}-Γj\Gamma_{j}-module in the sense of Definition 2.2.

Moreover, the fact that the Γj\Gamma_{j}-action on MjM_{j} is free and proper implies that L2(Ej)L^{2}(E_{j}) is admissible in the sense of Definition 2.3. To see this, choose a compact, Borel fundamental domain 𝒟j\mathcal{D}_{j} for the Γj\Gamma_{j}-action on MjM_{j} such that for each γΓj\gamma\in\Gamma_{j}, the restriction

pj|γ𝒟j:γ𝒟jNp_{j}|_{\gamma\cdot\mathcal{D}_{j}}\colon\gamma\cdot\mathcal{D}_{j}\rightarrow N

is a Borel isomorphism, where the projection maps pjp_{j} are as in subsection 2.3. For each j=1,2j=1,2 we have a map

Φj:L2(Ej)\displaystyle\Phi_{j}\colon L^{2}(E_{j}) l2(Γj)L2(Ej|𝒟j),\displaystyle\to l^{2}(\Gamma_{j})\otimes L^{2}(E_{j}|_{\mathcal{D}_{j}}),
s\displaystyle s γΓjγγ1χγ𝒟js.\displaystyle\mapsto\sum_{\gamma\in\Gamma_{j}}\gamma\otimes\gamma^{-1}\chi_{\gamma\mathcal{D}_{j}}s. (2.1)

This is a Γj\Gamma_{j}-equivariant unitary isomorphism with respect to the tensor product of the left-regular representation on l2(Γj)l^{2}(\Gamma_{j}) and the trivial representation on L2(Ej|𝒟j)L^{2}(E_{j}|_{\mathcal{D}_{j}}). Conjugation by Φj\Phi_{j} induces a *-isomorphism

[Mj]ΓjΓj𝒦(L2(Ej|𝒟j)).\mathbb{C}[M_{j}]^{\Gamma_{j}}\cong\mathbb{C}\Gamma_{j}\otimes\mathcal{K}(L^{2}(E_{j}|_{\mathcal{D}_{j}})). (2.2)
Remark 2.9.

It follows from (2.2) that for any r>0r>0, there exists a constant CrC_{r} such that for all T[Mj]ΓjT\in\mathbb{C}[M_{j}]^{\Gamma_{j}} with prop(T)r\textnormal{prop}(T)\leq r, we have

TmaxCrT(L2(Ej)).\|T\|_{\textnormal{max}}\leq C_{r}\|T\|_{\mathcal{B}(L^{2}(E_{j}))}.

Thus the maximal equivariant Roe algebra Cmax(Mj)ΓjC^{*}_{\textnormal{max}}(M_{j})^{\Gamma_{j}} from Definition 2.5 is well-defined in our setting. Moreover,

Cmax(Mj)ΓjCmax(Γj)𝒦,C^{*}_{\textnormal{max}}(M_{j})^{\Gamma_{j}}\cong C^{*}_{\textnormal{max}}(\Gamma_{j})\otimes\mathcal{K},

so that the KK-theories of both sides are isomorphic.


3. Folding maps

In this section, we define certain natural *-homomorphisms, called folding maps, between geometric CC^{*}-algebras of covering spaces, and discuss the role they play in functoriality for higher invariants. These maps were introduced in [3, Lemma 2.12].

To begin, let us provide some motivation at the level of groups. Observe that the quotient homomorphism Γ1Γ2Γ1/H\Gamma_{1}\rightarrow\Gamma_{2}\cong\Gamma_{1}/H induces a natural surjective *-homomorphism between group algebras,

α:Γ1Γ2,i=1kcγiγii=1kcγi[γi],\alpha\colon\mathbb{C}\Gamma_{1}\rightarrow\mathbb{C}\Gamma_{2},\quad\sum_{i=1}^{k}c_{\gamma_{i}}\gamma_{i}\mapsto\sum_{i=1}^{k}c_{\gamma_{i}}[\gamma_{i}],

where [γ][\gamma] is the class of an element γΓ1\gamma\in\Gamma_{1} in Γ1/H\Gamma_{1}/H. If one views elements of Γj\mathbb{C}\Gamma_{j} as kernel operators on the Hilbert space l2(Γj)l^{2}(\Gamma_{j}), then the map α\alpha takes a kernel k:Γ1×Γ1k\colon\Gamma_{1}\times\Gamma_{1}\rightarrow\mathbb{C} to the kernel

α(k):Γ2×Γ2,([γ],[γ])hHk(hγ,γ).\alpha(k)\colon\Gamma_{2}\times\Gamma_{2}\rightarrow\mathbb{C},\quad\big{(}[\gamma],[\gamma^{\prime}]\big{)}\mapsto\sum_{h\in H}k(h\gamma,\gamma^{\prime}).

There is an analogous map between kernels at the level of the Galois covers M1M_{1} and M2M1/HM_{2}\cong M_{1}/H. Given a smooth, Γ1\Gamma_{1}-equivariant Schwartz kernel k(x,y)k(x,y) with finite propagation on M1M_{1}, one can define a smooth, Γ2\Gamma_{2}-equivariant Schwartz kernel ψ(k)\psi(k) with finite propagation on M2M_{2} by the formula

ψ(k)([x],[y])=hHk(hx,y).\psi(k)([x],[y])=\sum_{h\in H}k(hx,y). (3.1)

Note that this sum is finite by properness of the HH-action on M1M_{1}. The formula (3.1) defines a map

ψ:𝒮Γ1𝒮Γ2,\psi\colon\mathcal{S}^{\Gamma_{1}}\to\mathcal{S}^{\Gamma_{2}},

where the kernel algebras 𝒮Γ1\mathcal{S}^{\Gamma_{1}} and 𝒮Γ2\mathcal{S}^{\Gamma_{2}} are as in Remark 2.7. This map is a *-homomorphism by [9, Lemma 5.2].

3.1. Definition of the folding map

We now define a more general version of the map (3.1), called the folding map Ψ\Psi, at the level of finite-propagation operators.

Let fp(L2(Ej))Γj\mathcal{B}_{\textnormal{fp}}(L^{2}(E_{j}))^{\Gamma_{j}} denote the *-algebra of bounded, Γj\Gamma_{j}-equivariant operators on L2(Ej)L^{2}(E_{j}) with finite propagation. The folding map Ψ\Psi is a *-homomorphism fp(L2(E1))Γ1fp(L2(E2))Γ2\mathcal{B}_{\textnormal{fp}}(L^{2}(E_{1}))^{\Gamma_{1}}\to\mathcal{B}_{\textnormal{fp}}(L^{2}(E_{2}))^{\Gamma_{2}} with the following properties:

  • For any Tfp(L2(E1))Γ1T\in\mathcal{B}_{\textnormal{fp}}(L^{2}(E_{1}))^{\Gamma_{1}}, we have prop(Ψ(T))prop(T)\textnormal{prop}\big{(}\Psi(T)\big{)}\leq\textnormal{prop}(T);

  • Ψ\Psi is surjective;

  • Ψ\Psi restricts to a surjective *-homomorphism [M1]Γ1[M2]Γ2\mathbb{C}[M_{1}]^{\Gamma_{1}}\rightarrow\mathbb{C}[M_{2}]^{\Gamma_{2}}.

These properties are proved in Propositions 3.3, 3.4, and 3.5 below.

To define Ψ\Psi, it will be convenient to use partitions of unity on M1M_{1} and M2M_{2} that are compatible with the Γ1\Gamma_{1} and Γ2\Gamma_{2}-actions, as follows. Since NN is compact, there exists ϵ>0\epsilon>0 such that for each xNx\in N, the ball Bϵ(x)B_{\epsilon}(x) is evenly covered with respect to both p1:M1Np_{1}\colon M_{1}\to N and p2:M2Np_{2}\colon M_{2}\to N. Let

𝒰N{Ui}iI\mathcal{U}_{N}\coloneqq\{U_{i}\}_{i\in I} (3.2)

be a finite open cover of NN such that each UiU_{i} has diameter at most ϵ\epsilon. Let {ϕi}iI\{\phi_{i}\}_{i\in I} be a partition of unity subordinate to 𝒰N\mathcal{U}_{N}.

For each iIi\in I, let UieU_{i}^{e} be a lift of UiU_{i} to M1M_{1} via the covering map p1:M1Np_{1}\colon M_{1}\to N. Similarly, let ϕie\phi_{i}^{e} be the lift of ϕi\phi_{i} to UieU_{i}^{e}. For each gΓ1g\in\Gamma_{1}, let UigU_{i}^{g} and ϕig\phi_{i}^{g} be the gg-translates of UieU_{i}^{e} and ϕie\phi_{i}^{e} in M1M_{1} respectively. Then

𝒰M1{Uig}iI,gΓ1,{ϕig}iI,gΓ1\mathcal{U}_{M_{1}}\coloneqq\big{\{}U_{i}^{g}\big{\}}_{\begin{subarray}{c}i\in I,\,g\in\Gamma_{1}\end{subarray}},\qquad\{\phi_{i}^{g}\}_{i\in I,\,g\in\Gamma_{1}} (3.3)

define a locally finite, Γ1\Gamma_{1}-invariant open cover of M1M_{1} and a subordinate partition of unity. This partition of unity is Γ1\Gamma_{1}-equivariant in the sense that for each iIi\in I, gΓg\in\Gamma, and xsupp(ϕie)x\in\textnormal{supp}(\phi_{i}^{e}), we have ϕig(gx)=ϕie(x)\phi_{i}^{g}(gx)=\phi_{i}^{e}(x).

After taking a quotient by the HH-action, we obtain a Γ2\Gamma_{2}-invariant open cover of M2M_{2}, together with a subordinate Γ2\Gamma_{2}-equivariant partition of unity:

𝒰M2{Ui[g]}iI,[g]Γ2,{ϕi[g]}iI,[g]Γ2.\mathcal{U}_{M_{2}}\coloneqq\big{\{}U_{i}^{[g]}\big{\}}_{\begin{subarray}{c}i\in I,\,[g]\in\Gamma_{2}\end{subarray}},\qquad\big{\{}\phi_{i}^{[g]}\big{\}}_{i\in I,\,[g]\in\Gamma_{2}}. (3.4)

For each ii and gg, we have the following commutative diagram of isomorphisms of local structures involving the covering map π\pi:

E1|Uig{E_{1}|_{U_{i}^{g}}}Uig{U_{i}^{g}}E2|Ui[g]{E_{2}|_{U_{i}^{[g]}}}Ui[g].{\,U_{i}^{[g]}.}π\scriptstyle{\pi}

We will write π\pi_{*} and π|Uig\pi|_{U_{i}^{g}}^{*} for the maps relating a local section of E1E_{1} to the corresponding local section of E2E_{2}. That is, if uu is a section of E1E_{1} over UigU_{i}^{g}, and vv is the corresponding section of E2E_{2} over Ui[g]U_{i}^{[g]}, then we have

u{u}v{v}π\scriptstyle{\pi_{*}}π|Uig\scriptstyle{\pi|_{U_{i}^{g}}^{*}}

If ww is any vector in the bundle E1E_{1}, we will also write π(w)\pi_{*}(w) for its image under the projection E1E2E_{1}\to E_{2}.

We are now ready for the definition of the folding map Ψ\Psi. Choose a set SΓ1S\subseteq\Gamma_{1} of coset representatives for Γ2H\Γ1\Gamma_{2}\cong H\backslash\Gamma_{1}. Given an operator Tfp(L2(E1))Γ1T\in\mathcal{B}_{\textnormal{fp}}(L^{2}(E_{1}))^{\Gamma_{1}}, we define Ψ(T)\Psi(T) to be the operator on L2(E2)L^{2}(E_{2}) that takes a section uL2(E2)u\in L^{2}(E_{2}) to the section

Ψ(T)ugΓ1,sSi,jIπ(ϕigTπ|Ujsϕj[s](u))L2(E2).\displaystyle\Psi(T)u\coloneqq\sum_{\begin{subarray}{c}g\in\Gamma_{1},\\ s\in S\end{subarray}}\sum_{i,j\in I}\pi_{*}\big{(}\phi_{i}^{g}\cdot T\circ\pi|_{U_{j}^{s}}^{*}\circ\phi_{j}^{[s]}(u)\big{)}\in L^{2}(E_{2}). (3.5)

To clarify the presentation of this equation and others like it, we adopt the following notational convention for moving between equivalent local sections of E1E_{1} and E2E_{2}:

Convention 3.1.

For vL2(E1)v\in L^{2}(E_{1}) and uL2(E2)u\in L^{2}(E_{2}), we will use the short-hand

  • ϕi[g]v\phi_{i}^{[g]}v to denote π(ϕigv)\pi_{*}(\phi_{i}^{g}v);

  • ϕigu\phi_{i}^{g}u to denote π|Uig(ϕi[g]u)\pi|^{*}_{U_{i}^{g}}(\phi_{i}^{[g]}u).

(Note that the meaning of ϕi[g]v\phi_{i}^{[g]}v depends on the representative gg.)

Using this convention, the formula (3.5) reads

Ψ(T)ugΓ1,sSi,jIϕi[g]T(ϕjsu)L2(E2).\displaystyle\Psi(T)u\coloneqq\sum_{\begin{subarray}{c}g\in\Gamma_{1},\\ s\in S\end{subarray}}\sum_{i,j\in I}\phi_{i}^{[g]}T(\phi_{j}^{s}u)\in L^{2}(E_{2}). (3.6)

We call Ψ\Psi the folding map. The fact that the above definition is independent of the choices of the set SS and compatible partitions of unity is proved in Proposition 3.3.

Before proceeding further, let us record a few straightfoward identities that will help us navigate through notational clutter:

  1. (i)

    If vv is a local section of E1E_{1} supported in some neighborhood UjgU_{j}^{g}, then for any gΓ1g^{\prime}\in\Gamma_{1} we have

    g(ϕigv)\displaystyle g^{\prime}\big{(}\phi_{i}^{g}v\big{)} =ϕigggv,\displaystyle=\phi_{i}^{g^{\prime}g}g^{\prime}v, (3.7)
    π(gv)\displaystyle\pi_{*}(g^{\prime}v) =[g]π(v).\displaystyle=[g^{\prime}]\cdot\pi_{*}(v). (3.8)
  2. (ii)

    If uu is a local section of E2E_{2} supported in some neighborhood Uj[g]U_{j}^{[g]}, then for any g,gΓ1g,g^{\prime}\in\Gamma_{1} and jIj\in I we have

    g(π|Ujgu)\displaystyle g^{\prime}\big{(}\pi|_{U_{j}^{g}}^{*}u\big{)} =π|Ujgg([g]u),\displaystyle=\pi|_{U_{j}^{g^{\prime}g}}^{*}([g^{\prime}]\cdot u), (3.9)
    u\displaystyle u =iI,γΓ1ϕi[γ](π|Ujgu).\displaystyle=\sum_{\begin{subarray}{c}i\in I,\\ \gamma\in\Gamma_{1}\end{subarray}}\phi_{i}^{[\gamma]}(\pi|_{U_{j}^{g}}^{*}u). (3.10)

The following lemma provides a convenient pointwise formula for Ψ(T)u\Psi(T)u:

Lemma 3.2.

For any uL2(E2)u\in L^{2}(E_{2}) and xM2x\in M_{2}, we have

(Ψ(T)u)(x)=π(jI,gΓ1T(ϕjgu)(y0)),\big{(}\Psi(T)u)(x)=\pi_{*}\Big{(}\sum_{\begin{subarray}{c}j\in I,\\ g\in\Gamma_{1}\end{subarray}}T(\phi_{j}^{g}u)(y_{0})\Big{)}, (3.11)

where y0y_{0} is any point in π1(x)\pi^{-1}(x), and the sum on the right-hand side is finite.

Proof.

Observe that for any jIj\in I and gΓ1g\in\Gamma_{1}, the summand T(ϕjgu)(y0)T(\phi_{j}^{g}u)(y_{0}) may only be non-zero if UjgBpropT(y0)U_{j}^{g}\cap B_{\textnormal{prop}\,T}(y_{0})\neq\emptyset. Since TT has finite propagation and the Γ1\Gamma_{1}-action is proper, the set

{(g,j)|UjgBpropT(y0)}\{(g,j)\,|\,U_{j}^{g}\cap B_{\textnormal{prop}\,T}(y_{0})\neq\emptyset\}

is finite. Hence the sum in question is finite for every xM2x\in M_{2} and y0π1(x)y_{0}\in\pi^{-1}(x).

To prove (3.11), let us first rewrite the right-hand side as

π(jIhH,sST(ϕjhsu)(y0)).\pi_{*}\Big{(}\sum_{\begin{subarray}{c}j\in I\end{subarray}}\sum_{\begin{subarray}{c}h\in H,\\ s\in S\end{subarray}}T(\phi_{j}^{hs}u)(y_{0})\Big{)}. (3.12)

Next, using (3.9), (3.10), and Γ1\Gamma_{1}-equivariance of TT, each summand in (3.12) equals

T(ϕjhsu)(y0)\displaystyle T(\phi_{j}^{hs}u)(y_{0}) =T(h(ϕjs([h]u)))(y0)\displaystyle=T\big{(}h(\phi_{j}^{s}([h]\cdot u))\big{)}(y_{0})
=T(h(ϕjsu))(y0)\displaystyle=T\big{(}h(\phi_{j}^{s}u)\big{)}(y_{0})
=h(T(ϕjsu))(y0)\displaystyle=h\big{(}T(\phi_{j}^{s}u)\big{)}(y_{0})
=(hiI,gΓ1ϕigTϕjsu)(y0),\displaystyle=\Big{(}h\cdot\sum_{\begin{subarray}{c}i\in I,\\ g\in\Gamma_{1}\end{subarray}}\phi_{i}^{g}T\phi_{j}^{s}u\Big{)}(y_{0}), (3.13)

where we have used Convention 3.1 when writing ϕjsu\phi_{j}^{s}u. Now observe that if vv is a section of E1E_{1} supported in some open set UigU_{i}^{g}, then for any xM2x\in M_{2}, we have

(πv)(x)=π(hH(hv)(y0)),(\pi_{*}v)(x)=\pi_{*}\big{(}\sum_{h\in H}(hv)(y_{0})\big{)}, (3.14)

where y0y_{0} is an arbitrary point in the inverse image π1(x)\pi^{-1}(x). Keeping this in mind while summing (3.1) over jIj\in I, sSs\in S, and hHh\in H, we see that (3.12) equals

π(h,g,s,i,j(h(ϕigTϕjsu))(y0))=(πg,s,i,jϕigT(ϕjsu))(x),\displaystyle\pi_{*}\Big{(}\sum_{h,g,s,i,j}\big{(}h(\phi_{i}^{g}T\phi_{j}^{s}u)\big{)}(y_{0})\Big{)}=\Big{(}\pi_{*}\sum_{g,s,i,j}\phi_{i}^{g}\cdot T(\phi_{j}^{s}u)\Big{)}(x),

which is equal to (Ψ(T)u)(x)(\Psi(T)u)(x) by (3.6) and noting Convention 3.1. ∎

Proposition 3.3.

The folding map Ψ\Psi defined in (3.6) is a *-homomorphism

Ψ:fp(L2(E1))Γ1\displaystyle\Psi\colon\mathcal{B}_{\textnormal{fp}}(L^{2}(E_{1}))^{\Gamma_{1}} fp(L2(E2))Γ2\displaystyle\rightarrow\mathcal{B}_{\textnormal{fp}}(L^{2}(E_{2}))^{\Gamma_{2}}

and is independent of the choices of the set SS of coset representatives and compatible partitions of unity used to define it.

Proof.

Boundedness of Ψ(T)\Psi(T) follows from boundedness and finite propagation of TT, via the formula (3.11). To see that Ψ(T)\Psi(T) is Γ2\Gamma_{2}-equivariant, note that by (3.11), we have for all [γ]Γ2[\gamma]\in\Gamma_{2} and uL2(E2)u\in L^{2}(E_{2}) that

([γ]Ψ(T)[γ1]u)(x)\displaystyle([\gamma]\Psi(T)[\gamma^{-1}]u)(x) =[γ]πjI,gΓ1T(ϕjg[γ1]u)(γ1y0)\displaystyle=[\gamma]\cdot\pi_{*}\sum_{\begin{subarray}{c}j\in I,\\ g\in\Gamma_{1}\end{subarray}}T(\phi_{j}^{g}[\gamma^{-1}]u)(\gamma^{-1}y_{0})
=[γ]πj,gTγ1(ϕjγgu)(γ1y0).\displaystyle=[\gamma]\cdot\pi_{*}\sum_{\begin{subarray}{c}j,g\end{subarray}}T\gamma^{-1}(\phi_{j}^{\gamma g}u)(\gamma^{-1}y_{0}).

Using (3.8) and Γ1\Gamma_{1}-equivariance of TT, this is equal to

πj,gγTγ1(ϕjγgu)(y0)\displaystyle\pi_{*}\sum_{j,g}\gamma T\gamma^{-1}(\phi_{j}^{\gamma g}u)(y_{0}) =πj,gT(ϕjγgu)(y0)=(Ψ(T)u)(x),\displaystyle=\pi_{*}\sum_{j,g}T(\phi_{j}^{\gamma g}u)(y_{0})=(\Psi(T)u)(x),

where we used a change-of-variable for the last equality.

To see that Ψ\Psi is a *-homomorphism, let QQ be another operator in fp(L2(E1))Γ1\mathcal{B}_{\textnormal{fp}}(L^{2}(E_{1}))^{\Gamma_{1}}. Let xM2x\in M_{2} and y0π1(x)y_{0}\in\pi^{-1}(x). By (3.11), and the fact that both TT^{\prime} and TT have finite propagation, one sees that there exists a finite subset FΓ1F\subseteq\Gamma_{1} such that

(Ψ(TT)u)(x)\displaystyle(\Psi(T^{\prime}T)u)(x) =πjI,gFTT(ϕjgu)(y0)\displaystyle=\pi_{*}\sum_{\begin{subarray}{c}j\in I,\\ g\in F\end{subarray}}T^{\prime}T(\phi_{j}^{g}u)(y_{0})
=π(Tj,gT(ϕjgu))(y0)\displaystyle=\pi_{*}\Big{(}T^{\prime}\sum_{j,g}T(\phi_{j}^{g}u)\Big{)}(y_{0})
=π(TiI,γFϕiγj,gT(ϕjgu))(y0)\displaystyle=\pi_{*}\Big{(}T^{\prime}\sum_{\begin{subarray}{c}i\in I,\\ \gamma\in F\end{subarray}}\phi_{i}^{\gamma}\sum_{j,g}T(\phi_{j}^{g}u)\Big{)}(y_{0})
=π(i,γT(ϕiγj,gT(ϕjgu)))(y0)\displaystyle=\pi_{*}\Big{(}\sum_{i,\gamma}T^{\prime}\big{(}\phi_{i}^{\gamma}\sum_{j,g}T(\phi_{j}^{g}u)\big{)}\Big{)}(y_{0})
=(Ψ(T)Ψ(T)u)(x).\displaystyle=(\Psi(T^{\prime})\circ\Psi(T)u)(x).

One checks directly that Ψ\Psi respects *-operations.

That (3.6) is independent of the choice of coset representatives SΓ1S\subseteq\Gamma_{1} is implied by the pointwise formula (3.11). To see that (3.6) is independent of partitions of unity, let {φi}\{\varphi_{i}\}, {φi[γ]}\{\varphi_{i}^{[\gamma]}\}, and {φiγ}\{\varphi_{i}^{\gamma}\} be another set of compatible partitions of unity for NN, M2M_{2}, and M1M_{1} with the same properties as {ϕi}\{\phi_{i}\}, {ϕi[γ]}\{\phi_{i}^{[\gamma]}\}, and {ϕiγ}\{\phi_{i}^{\gamma}\}. Let us write

Ψ{ϕ}(T)andΨ{φ}(T)\Psi^{\{\phi\}}(T)\quad\textnormal{and}\quad\Psi^{\{\varphi\}}(T)

to distinguish the operators defined by (3.6) with respect to these two sets of partitions of unity. Since TT has finite propagation, there exists a finite subset FΓ1F^{\prime}\subseteq\Gamma_{1} such that if gFg\notin F, then supp(ϕjg)BpropT(y0)=\operatorname{supp}(\phi_{j}^{g})\cap B_{\textnormal{prop}\,T}(y_{0})=\emptyset and supp(φjg)BpropT(y0)=\operatorname{supp}(\varphi_{j}^{g})\cap B_{\textnormal{prop}\,T}(y_{0})=\emptyset for any jIj\in I. By (3.11), for each xM2x\in M_{2} we have

(Ψ{ϕ}(T)u)(x)\displaystyle(\Psi^{\{\phi\}}(T)u)(x) =πjI,gΓ1T(ϕjgu)(y0)\displaystyle=\pi_{*}\sum_{j\in I,\,g\in\Gamma_{1}}T(\phi_{j}^{g}u)(y_{0})
=πjI,gFT(𝟙BpropT(y0)ϕjgu)(y0)\displaystyle=\pi_{*}\sum_{j\in I,\,g\in F}T(\mathbbm{1}_{B_{\textnormal{prop}\,T}(y_{0})}\phi_{j}^{g}u)(y_{0})
=π((Tj,g𝟙BpropT(y0)ϕjgu)(y0)),\displaystyle=\pi_{*}\Big{(}\big{(}T\sum_{\begin{subarray}{c}j,g\end{subarray}}\mathbbm{1}_{B_{\textnormal{prop}\,T}(y_{0})}\phi_{j}^{g}u\big{)}(y_{0})\Big{)}, (3.15)

where 𝟙BpropT(y0)\mathbbm{1}_{B_{\textnormal{prop}\,T}(y_{0})} is the characteristic function of the set BpropT(y0)B_{\textnormal{prop}\,T}(y_{0}), and we have used that TT commutes with finite sums. Likewise, we have

(Ψ{φ}(T)u)(x)\displaystyle(\Psi^{\{\varphi\}}(T)u)(x) =π((Tj,g𝟙BpropT(y0)φjgu)(y0)).\displaystyle=\pi_{*}\Big{(}\big{(}T\sum_{\begin{subarray}{c}j,g\end{subarray}}\mathbbm{1}_{B_{\textnormal{prop}\,T}(y_{0})}\varphi_{j}^{g}u\big{)}(y_{0})\Big{)}. (3.16)

Now observe that

j,g𝟙BpropT(y0)ϕjguandj,g𝟙BpropT(y0)φjgu\sum_{j,g}\mathbbm{1}_{B_{\textnormal{prop}\,T}(y_{0})}\phi_{j}^{g}u\quad\textnormal{and}\quad\sum_{j,g}\mathbbm{1}_{B_{\textnormal{prop}\,T}(y_{0})}\varphi_{j}^{g}u

are both equal to the lift of the section uu to M1M_{1} restricted to the compact subset BpropT(y0)B_{\textnormal{prop}\,T}(y_{0}). Thus the right-hand sides of (3.1) and (3.16) are equal, whence

(Ψ{ϕ}(T)(u))(x)=(Ψ{φ}(T)(u))(x).\displaystyle(\Psi^{\{\phi\}}(T)(u))(x)=(\Psi^{\{\varphi\}}(T)(u))(x).

The next proposition shows that the folding map Ψ\Psi preserves locality of operators. In particular, this means that Ψ\Psi induces a map at the level of localization algebras (see Definition 3.8), which is a crucial ingredient for our main result, Theorem 1.1 .

Proposition 3.4.

For any operator Tfp(L2(E1))Γ1T\in\mathcal{B}_{\textnormal{fp}}(L^{2}(E_{1}))^{\Gamma_{1}}, we have

prop(Ψ(T))prop(T).\textnormal{prop}\big{(}\Psi(T)\big{)}\leq\textnormal{prop}(T).
Proof.

Take a section uCc(E2)u\in C_{c}(E_{2}), and write it as a finite sum j,[g]ϕj[g]u\sum_{j,[g]}\phi_{j}^{[g]}u. By (3.11),

(Ψ(T)u)(x)=πjI,gΓ1T(ϕjgu)(y0),\big{(}\Psi(T)u)(x)=\pi_{*}\sum_{j\in I,g\in\Gamma_{1}}T(\phi_{j}^{g}u)(y_{0}),

where y0y_{0} is a lift of xx to M1M_{1}. Now if dM2(x,suppu)>prop(T)d_{M_{2}}(x,\operatorname{supp}u)>\operatorname{prop}(T), then

dM1(y0,supp(ϕjgu))>prop(T)d_{M_{1}}\big{(}y_{0},\operatorname{supp}(\phi_{j}^{g}u)\big{)}>\operatorname{prop}(T)

for any jIj\in I and gΓ1g\in\Gamma_{1}, since distances between points do not increase under the projection π\pi. Thus if dM2(x,suppu)>prop(T)d_{M_{2}}(x,\operatorname{supp}u)>\operatorname{prop}(T), then Ψ(T)u(x)=0\Psi(T)u(x)=0. Since this holds for any section uCc(E2)u\in C_{c}(E_{2}), we have prop(Ψ(T))prop(T)\operatorname{prop}(\Psi(T))\leq\operatorname{prop}(T). ∎

Proposition 3.5.

The map Ψ\Psi is surjective and preserves local compactness of operators. Moreover, it restricts to a surjective *-homomorphism

Ψ:[M1]Γ1[M2]Γ2.\Psi\colon\mathbb{C}[M_{1}]^{\Gamma_{1}}\rightarrow\mathbb{C}[M_{2}]^{\Gamma_{2}}.
Proof.

Let SΓ1S\subseteq\Gamma_{1} be the set of coset representatives used in the definition of Ψ\Psi. For any T2fp(L2(E2))Γ2T_{2}\in\mathcal{B}_{\textnormal{fp}}(L^{2}(E_{2}))^{\Gamma_{2}}, define an operator T1T_{1} on L2(E1)L^{2}(E_{1}) by the formula

T1(v)gΓ1iIsSjIϕjgsT2(ϕi[g]v),T_{1}(v)\coloneqq\sum_{\begin{subarray}{c}g\in\Gamma_{1}\\ i\in I\,\,\end{subarray}}\sum_{\begin{subarray}{c}s\in S\\ j\in I\end{subarray}}\phi_{j}^{gs}T_{2}\big{(}\phi_{i}^{[g]}v\big{)}, (3.17)

for vL2(E1)v\in L^{2}(E_{1}). We claim that T1fp(L2(E1))Γ1T_{1}\in\mathcal{B}_{\textnormal{fp}}(L^{2}(E_{1}))^{\Gamma_{1}} and that Ψ(T1)=T2\Psi(T_{1})=T_{2}. Indeed, for each fixed ii and jj, the sum of operators

gΓ1sSϕjgsT2ϕi[g]\sum_{\begin{subarray}{c}g\in\Gamma_{1}\end{subarray}}\sum_{\begin{subarray}{c}s\in S\end{subarray}}\phi_{j}^{gs}T_{2}\phi_{i}^{[g]}

converges strongly in (L2(E1))\mathcal{B}(L^{2}(E_{1})). Since II is a finite indexing set, it follows that T1T_{1} is bounded. To see that T1T_{1} is Γ1\Gamma_{1}-equivariant, note that for any γΓ1\gamma\in\Gamma_{1}, we have, by the identities (3.7) and (3.8), that

γ1T1γ(v)\displaystyle\gamma^{-1}T_{1}\gamma(v) =g,is,jγ1(ϕjgsT2(ϕi[g]γv))\displaystyle=\sum_{g,i}\sum_{s,j}\gamma^{-1}\big{(}\phi_{j}^{gs}T_{2}(\phi_{i}^{[g]}\gamma v)\big{)}
=g,is,jγ1(ϕjgsT2(π(γ(ϕiγ1gv))))\displaystyle=\sum_{g,i}\sum_{s,j}\gamma^{-1}\big{(}\phi_{j}^{gs}T_{2}(\pi_{*}(\gamma\cdot(\phi_{i}^{\gamma^{-1}g}v)))\big{)}
=g,is,jγ1(ϕjgsT2([γ](ϕi[γ1g]v))),\displaystyle=\sum_{g,i}\sum_{s,j}\gamma^{-1}\big{(}\phi_{j}^{gs}T_{2}([\gamma]\cdot(\phi_{i}^{[\gamma^{-1}g]}v))\big{)},

where we observe Convention 3.1 as usual. Using Γ2\Gamma_{2}-equivariance of T2T_{2} and (3.9), one finds this to be equal to

g,is,jγ1(ϕj[gs]([γ]T2(ϕi[γ1g]v)))\displaystyle\sum_{g,i}\sum_{s,j}\gamma^{-1}\big{(}\phi_{j}^{[gs]}([\gamma]T_{2}(\phi_{i}^{[\gamma^{-1}g]}v))\big{)}
=\displaystyle= g,is,jϕjγ1gsT2(ϕiγ1gv)\displaystyle\sum_{g,i}\sum_{s,j}\phi_{j}^{\gamma^{-1}gs}T_{2}(\phi_{i}^{\gamma^{-1}g}v)
=\displaystyle= γ1g,is,jϕjγ1gsT2(ϕi[γ1g]v)\displaystyle\sum_{\gamma^{-1}g,i}\sum_{s,j}\phi_{j}^{\gamma^{-1}gs}T_{2}(\phi_{i}^{[\gamma^{-1}g]}v)
=\displaystyle= T1(v),\displaystyle\,\,T_{1}(v),

using the definition of T1T_{1} and a change of variable for the last equality.

Next, finite propagation of T1T_{1} follows from finite propagation of T2T_{2}. Indeed, for any vCc(E2)v\in C_{c}(E_{2}), gΓ1g\in\Gamma_{1}, and iIi\in I, the set

ST1{sS|ϕj[gs]T2(ϕi[g]v)0 for some jI}S_{T_{1}}\coloneqq\left\{s\in S\,\big{|}\,\phi_{j}^{[gs]}T_{2}\big{(}\phi_{i}^{[g]}v\big{)}\neq 0\textnormal{ for some }j\in I\right\}

is finite, thus the sum over SS in equation (3.17) reduces to a finite sum over ST1S_{T_{1}}. By the same equation, and the fact that the Γ1\Gamma_{1}-action is isometric, we have

prop(T1)\displaystyle\textnormal{prop}(T_{1}) supsST1i,jI{dM1(Uig,Ujgs)}=supsST1i,jI{dM1(Uie,Ujs)}<.\displaystyle\leq\sup_{\begin{subarray}{c}\,\,s\in S_{T_{1}}\\ i,j\in I\end{subarray}}\{d_{M_{1}}\big{(}U_{i}^{g},U_{j}^{gs}\big{)}\}=\sup_{\begin{subarray}{c}\,\,s\in S_{T_{1}}\\ i,j\in I\end{subarray}}\left\{d_{M_{1}}\big{(}U_{i}^{e},U_{j}^{s}\big{)}\right\}<\infty.

We now show that Ψ(T1)=T2\Psi(T_{1})=T_{2}. By (3.17) and (3.5) we have, for any uCc(E2)u\in C_{c}(E_{2}),

Ψ(T1)u\displaystyle\Psi(T_{1})u =γΓ1,tSi,jIϕi[γ]T1ϕjt(u)\displaystyle=\sum_{\begin{subarray}{c}\gamma\in\Gamma_{1},\\ t\in S\end{subarray}}\sum_{i,j\in I}\phi_{i}^{[\gamma]}T_{1}\phi_{j}^{t}(u)
=γΓ1,tSi,jIϕi[γ](gΓ1kIsSlIϕlgsT2(ϕk[g](ϕjtu))).\displaystyle=\sum_{\begin{subarray}{c}\gamma\in\Gamma_{1},\\ t\in S\end{subarray}}\sum_{i,j\in I}\phi_{i}^{[\gamma]}\Big{(}\sum_{\begin{subarray}{c}g\in\Gamma_{1}\\ k\in I\,\,\end{subarray}}\sum_{\begin{subarray}{c}s\in S\\ l\in I\end{subarray}}\phi_{l}^{gs}T_{2}\big{(}\phi_{k}^{[g]}(\phi_{j}^{t}u)\big{)}\Big{)}.

Finite propagation of T2T_{2} and compact support of uu imply that all of these sums are finite, so by (3.10) this is equal to

g,s,k,lγ,t,i,jϕi[γ](ϕlgsT2(ϕk[g](ϕjtu)))=g,s,k,lt,jϕl[gs]T2(ϕk[g](ϕjtu)).\displaystyle\sum_{g,s,k,l}\sum_{\gamma,t,i,j}\phi_{i}^{[\gamma]}\big{(}\phi_{l}^{gs}T_{2}\big{(}\phi_{k}^{[g]}(\phi_{j}^{t}u)\big{)}\big{)}=\sum_{g,s,k,l}\sum_{t,j}\phi_{l}^{[gs]}T_{2}\big{(}\phi_{k}^{[g]}(\phi_{j}^{t}u)\big{)}.

Since {gs|sS}\{gs\,|\,s\in S\} is a set of coset representatives for H\Γ1H\backslash\Gamma_{1} for any gΓ1g\in\Gamma_{1}, the identity (3.10) implies that the above is equal to

t,jT2(g,kϕk[g](ϕjtu))=t,jT2(ϕj[t]u)=T2u.\displaystyle\sum_{t,j}T_{2}\Big{(}\sum_{g,k}\phi_{k}^{[g]}(\phi_{j}^{t}u)\Big{)}=\sum_{t,j}T_{2}(\phi_{j}^{[t]}u)=T_{2}u.

Thus Ψ(T1)=T2\Psi(T_{1})=T_{2} as bounded operators on L2(E2)L^{2}(E_{2}).

Finally, that Ψ\Psi preserves local compactness of operators follows directly from (3.11), so that Ψ\Psi restricts to a map [M1]Γ1[M2]Γ2\mathbb{C}[M_{1}]^{\Gamma_{1}}\rightarrow\mathbb{C}[M_{2}]^{\Gamma_{2}}. To see that this map is surjective, suppose T2T_{2} is locally compact. Then for any fCc(M1)f\in C_{c}(M_{1}), we have

fT1=fgΓ1sSi,jIϕjgsT2ϕi[g].fT_{1}=f\sum_{\begin{subarray}{c}g\in\Gamma_{1}\\ s\in S\,\,\end{subarray}}\sum_{\begin{subarray}{c}i,j\in I\end{subarray}}\phi_{j}^{gs}T_{2}\phi_{i}^{[g]}.

The sum over Γ1\Gamma_{1} reduces to a sum over the finite set

Ff{gΓ1|Ujgssupp(f) for some jI},F_{f}\coloneqq\left\{g\in\Gamma_{1}\,\big{|}\,U_{j}^{gs}\cap\operatorname{supp}(f)\neq\emptyset\textnormal{ for some }j\in I\right\},

hence fT1fT_{1} is equal to

gFfsSi,jIfϕjgs𝟙π(suppf)T2ϕi[g].\displaystyle\sum_{\begin{subarray}{c}g\in F_{f}\\ s\in S\,\,\end{subarray}}\sum_{\begin{subarray}{c}i,j\in I\end{subarray}}f\phi_{j}^{gs}\mathbbm{1}_{\pi_{*}(\operatorname{supp}f)}T_{2}\phi_{i}^{[g]}.

Since the subset π(suppf)M2\pi_{*}(\operatorname{supp}f)\subseteq M_{2} is compact, local compactness of T2T_{2} means that this is a finite sum of compact operators and hence compact. A similar argument shows that the operator T1fT_{1}f is compact. Thus the operator T1T_{1} is locally compact, so Ψ\Psi restricts to a surjective *-homomorphism [M1]Γ1[M2]Γ2\mathbb{C}[M_{1}]^{\Gamma_{1}}\rightarrow\mathbb{C}[M_{2}]^{\Gamma_{2}}. ∎

3.2. Description of the folding map on invariant sections

The folding map Ψ\Psi admits an equivalent description using NN-invariant sections of E1E_{1}. Such sections, while not in general square-integrable over M1M_{1}, form a space that is naturally isomorphic to L2(E2)Γ2L^{2}(E_{2})^{\Gamma_{2}}, as we now describe. This allows computations involving Ψ\Psi to be carried out entirely on M1M_{1}. Thus it may be a useful perspective for some applications. The discussion below slightly generalizes the averaging map from [10, subsection 5.2] applied in the context of discrete groups.

Let c:M1[0,1]c\colon M_{1}\rightarrow[0,1] be a function whose support has compact intersections with every HH-orbit, and such that for all xM1x\in M_{1},

hHc(hx)2=1.\sum_{h\in H}c(hx)^{2}=1.

Note that this sum is finite by properness of the Γ1\Gamma_{1}-action.

Let Ctc(E1)HC_{\operatorname{tc}}(E_{1})^{H} denote the space of HH-transversally compactly supported sections of E1E_{1}, defined as the space of continuous, HH-invariant sections of E1E_{1} whose supports have compact images in M2M1/HM_{2}\cong M_{1}/H under the quotient map. Let LT2(E1)HL^{2}_{T}(E_{1})^{H} denote the Hilbert space of HH-invariant, HH-transversally L2L^{2}-sections of E1E_{1}, defined as the completion of Ctc(E1)HC_{\operatorname{tc}}(E_{1})^{H} with respect to the inner product

(s1,s2)LT2(E1)N(cs1,cs2)L2(E1).(s_{1},s_{2})_{L^{2}_{T}(E_{1})^{N}}\coloneqq(cs_{1},cs_{2})_{L^{2}(E_{1})}.

Then one checks that:

Lemma 3.6.

The space LT2(E1)HL^{2}_{T}(E_{1})^{H} is naturally unitarily isomorphic to L2(E2)L^{2}(E_{2}) and is independent of the choice of cc.

For any Tfp(L2(E1))Γ1T\in\mathcal{B}_{\textnormal{fp}}(L^{2}(E_{1}))^{\Gamma_{1}}, the operator Ψ(T)\Psi(T) on L2(E2)L^{2}(E_{2}) from (3.6) can be described equivalently by its action on the isomorphic space LT2(E1)HL^{2}_{T}(E_{1})^{H} as follows. Take a partition of unity {ϕig}\{\phi_{i}^{g}\} of M1M_{1} as in (3.3). Let sLT2(E1)Hs\in L^{2}_{T}(E_{1})^{H} be an HH-transversally L2L^{2}-section of E1E_{1}. For any yM1y\in M_{1}, set

(Ψ(T)s)(y)iI,gΓ1(T(ϕigs)(y)).(\Psi(T)s)(y)\coloneqq\sum_{i\in I,g\in\Gamma_{1}}\left(T(\phi_{i}^{g}s\right)(y)). (3.18)

The fact that TT has finite propagation means that this pointwise sum is finite, and that c(Ψ(T)s)L2(E1)c(\Psi(T)s)\in L^{2}(E_{1}). Further, Ψ(T)s\Psi(T)s is an element of LT2(E1)HL^{2}_{T}(E_{1})^{H}, since for any hHh\in H and yM1y\in M_{1} we have

(h(Ψ(T)s))(y)=iI,gΓ1(Th(ϕigs)(y))=iI,gΓ1T(ϕihgs)(y)=(Ψ(T)s)(y),(h(\Psi(T)s))(y)=\sum_{i\in I,g\in\Gamma_{1}}\left(T\circ h(\phi_{i}^{g}s\right)(y))=\sum_{i\in I,g\in\Gamma_{1}}T(\phi_{i}^{hg}s)(y)=(\Psi(T)s)(y),

where we have used the equivariance properties of TT and ss. Finally, a direct comparison shows that this definition is equivalent with that given by equation (3.11).

3.3. Induced maps on geometric CC^{*}-algebras and KK-theory

By Proposition 3.5, the folding map Ψ\Psi restricts to a surjective *-homomorphism

Ψ:[M1]Γ1[M2]Γ2.\Psi\colon\mathbb{C}[M_{1}]^{\Gamma_{1}}\rightarrow\mathbb{C}[M_{2}]^{\Gamma_{2}}.

By the defining property of maximal completions of *-algebras, Ψ\Psi extends to a surjective *-homomorphism between maximal equivariant Roe algebras,

Ψ:Cmax(M1)Γ1Cmax(M2)Γ2.\Psi\colon C^{*}_{\textnormal{max}}(M_{1})^{\Gamma_{1}}\rightarrow C^{*}_{\textnormal{max}}(M_{2})^{\Gamma_{2}}.

In this paper we will make use of a number natural extensions of this map to other geometric CC^{*}-algebras, as well as the induced maps on KK-theory. We refer to these maps collectively as folding maps.

To begin, we have the following elementary lemma:

Lemma 3.7.

Let AA and BB be CC^{*}-algebras and let ϕ:AB\phi\colon A\rightarrow B be a surjective *-homomorphism. Then ϕ\phi extends to a *-homomorphism ϕ~:(A)(B)\widetilde{\phi}\colon\mathcal{M}(A)\rightarrow\mathcal{M}(B).

Proof.

Since ϕ\phi is surjective, we may define ϕ~\widetilde{\phi} by requiring

ϕ~(m)ϕ(a)=ϕ(ma),ϕ(a)ϕ~(m)=ϕ(am),\widetilde{\phi}(m)\phi(a)=\phi(ma),\quad\phi(a)\widetilde{\phi}(m)=\phi(am),

for m𝒜m\in\mathcal{A}. Concretely, by way of a faithful non-degenerate representation σ:B(H)\sigma\colon B\rightarrow\mathcal{B}(H) on some Hilbert space HH, we may identify BB with a subalgebra of (H)\mathcal{B}(H). Since ϕ\phi is surjective, the composition σϕ:A(H)\sigma\circ\phi\colon A\rightarrow\mathcal{B}(H) is a non-degenerate representation of AA. Thus the above formulas define a representation of (A)\mathcal{M}(A) with values in the idealizer of σ(B)\sigma(B), which one identifies with (B)\mathcal{M}(B). (See also [7] Lemma I.9.14.) ∎

By Lemma 3.7 and Proposition 3.5, the map Ψ\Psi extends to a *-homomorphism

(Cmax(M1)Γ1)(Cmax(M2)Γ2).\mathcal{M}\left(C^{*}_{\textnormal{max}}(M_{1})^{\Gamma_{1}}\right)\rightarrow\mathcal{M}\left(C^{*}_{\textnormal{max}}(M_{2})^{\Gamma_{2}}\right).

Viewing Cmax(Mj)ΓjC^{*}_{\textnormal{max}}(M_{j})^{\Gamma_{j}} as an ideal in \mathcal{M}, we will still denote this extended map by Ψ\Psi. This map will be essential when we apply the functional calculus for the maximal Roe algebra in sections 5 and 6.

Next, suppose we have a path r:[0,)Cmax(M1)Γ1r\colon[0,\infty)\to C^{*}_{\textnormal{max}}(M_{1})^{\Gamma_{1}} that satisfies

prop(r(t))0ast.\textnormal{prop}(r(t))\to 0\quad\textnormal{as}\quad t\to\infty.

Then by Proposition 3.4, the same is true of the path Ψr:[0,)Cmax(M1)Γ2\Psi\circ r\colon[0,\infty)\rightarrow C^{*}_{\textnormal{max}}(M_{1})^{\Gamma_{2}}. This allows the folding map to pass to the level of localization algebras:

Definition 3.8.

Define the map

ΨL:CL,max(M1)Γ1\displaystyle\Psi_{L}\colon C^{*}_{L,\textnormal{max}}(M_{1})^{\Gamma_{1}} CL,max(M2)Γ2\displaystyle\to C^{*}_{L,\textnormal{max}}(M_{2})^{\Gamma_{2}}
r\displaystyle r Ψr.\displaystyle\mapsto\Psi\circ r.

Then ΨL\Psi_{L} restricts to a *-homomorphism between obstruction algebras:

ΨL,0:CL,0,max(M1)Γ1CL,0,max(M2)Γ2.\Psi_{L,0}\colon C^{*}_{L,0,\textnormal{max}}(M_{1})^{\Gamma_{1}}\to C^{*}_{L,0,\textnormal{max}}(M_{2})^{\Gamma_{2}}.

Each of the above maps Ψ\Psi, ΨL\Psi_{L}, and ΨL,0\Psi_{L,0} induces a map at the level of KK-theory. In this paper, we will make use of two of them:

Ψ:K(Cmax(M1)Γ1)\displaystyle\Psi_{*}\colon K_{\bullet}\left(C^{*}_{\textnormal{max}}(M_{1})^{\Gamma_{1}}\right) K(Cmax(M2)Γ2),\displaystyle\rightarrow K_{\bullet}\left(C^{*}_{\textnormal{max}}(M_{2})^{\Gamma_{2}}\right),
(ΨL,0):K(CL,0,max(M1)Γ1)\displaystyle(\Psi_{L,0})_{*}\colon K_{\bullet}\left(C^{*}_{L,0,\textnormal{max}}(M_{1})^{\Gamma_{1}}\right) K(CL,0,max(M2)Γ2).\displaystyle\rightarrow K_{\bullet}\left(C^{*}_{L,0,\textnormal{max}}(M_{2})^{\Gamma_{2}}\right).
Remark 3.9.

In the recent paper [21], Schick and Seyedhosseini have constructed a slightly different completion of the equivariant algebraic Roe algebra, called the quotient completion, which still behaves functorially under maps between covering spaces. For details of the construction, see [21, section 3].


4. Functional calculus and the wave operator

We now develop the analytical properties of the wave operator on the maximal Roe algebra that will form the basis of our work in subsequent sections.

Let us begin by recalling the functional calculus for the maximal Roe algebra, which was developed in a more general setting in [11]. The discussion here is specialized to the cocompact setting.

Throughout this section we will work in the geometric situation in subsection 2.3. To simplify notation, in this subsection and the next we will write M,ΓM,\Gamma for either M1,Γ1M_{1},\Gamma_{1} or M2,Γ2M_{2},\Gamma_{2}, and D,ED,E for either D1,E1D_{1},E_{1} or D2,E2D_{2},E_{2}. In other words, Γ\Gamma acts freely and properly on MM with compact quotient, and DD is a Γ\Gamma-equivariant operator on the bundle EME\to M.

4.1. Functional calculus on the maximal Roe algebra

We shall view the CC^{*}-algebra Cmax(M)ΓC^{*}_{\text{max}}(M)^{\Gamma} as a right Hilbert module over itself. The inner product and right action on Cmax(M)ΓC^{*}_{\text{max}}(M)^{\Gamma} are defined naturally through multiplication: for a,bCmax(M)Γa,b\in C^{*}_{\text{max}}(M)^{\Gamma},

a,b=ab,ab=ab.\langle a,b\rangle=a^{*}b,\qquad a\cdot b=ab. (4.1)

The algebra of compact operators on this Hilbert module can be identified with Cmax(M)ΓC^{*}_{\text{max}}(M)^{\Gamma} via left multiplication. Similarly, the algebra of bounded adjointable operators can be identified with the multiplier algebra \mathcal{M} of Cmax(M)ΓC^{*}_{\text{max}}(M)^{\Gamma}.

The operator DD defines an unbounded operator on this Hilbert module in the following way. First note that DD acts on smooth sections of the external tensor product EEM×ME\boxtimes E^{*}\rightarrow M\times M by taking a section ss to the section DsDs defined by

(Ds)(x,y)=Dxs(x,y),(Ds)(x,y)=D_{x}s(x,y), (4.2)

where DxD_{x} denotes DD acting on the xx-variable.

Let 𝒮Γ\mathcal{S}^{\Gamma} be the *-subalgebra of [M]Γ\mathbb{C}[M]^{\Gamma} defined in Remark 2.7. That is, an element of 𝒮Γ\mathcal{S}^{\Gamma} is an operator TκT_{\kappa} given by a smooth kernel κCb(EE)\kappa\in C_{b}^{\infty}(E\boxtimes E^{*}) that:

  1. (i)

    is Γ\Gamma-equivariant with respect to the diagonal Γ\Gamma-action;

  2. (ii)

    has finite propagation, meaning that there exists a constant cκ0c_{\kappa}\geq 0 such that if d(x,y)>cκd(x,y)>c_{\kappa} then κ(x,y)=0\kappa(x,y)=0.

The operator DD acts on elements of 𝒮Γ\mathcal{S}^{\Gamma} by acting on the corresponding smooth kernels as in (4.2). In this way, DD becomes a densely defined operator on the Hilbert module Cmax(M)ΓC^{*}_{\text{max}}(M)^{\Gamma} that one verifies is symmetric with respect to the inner product in (4.1).

Theorem 4.1 ([11] Theorem 3.1).

There exists a real number μ0\mu\neq 0 such that the operators

D±μi:Cmax(M)ΓCmax(M)ΓD\pm\mu i\colon C^{*}_{\text{max}}(M)^{\Gamma}\rightarrow C^{*}_{\text{max}}(M)^{\Gamma}

have dense range.

Consequently, the operator DD on the Hilbert module Cmax(M)ΓC^{*}_{\text{max}}(M)^{\Gamma} is regular and essentially self-adjoint, and so admits a continuous functional calculus (see [15, Theorem 10.9] and [14, Proposition 16]):

Theorem 4.2.

For j=1j=1 or 22, there is a *-preserving linear map

π:C()\displaystyle\pi\colon C(\mathbb{R}) (Cmax(M)Γ),\displaystyle\rightarrow\mathcal{R}\left(C^{*}_{\text{max}}(M)^{\Gamma}\right),
f\displaystyle f f(D)π(f),\displaystyle\mapsto f(D)\coloneqq\pi(f),

where C()C(\mathbb{R}) denotes the continuous functions \mathbb{R}\to\mathbb{C}, and (Cmax(M)Γ)\mathcal{R}\left(C^{*}_{\text{max}}(M)^{\Gamma}\right) denotes the regular operators on Cmax(M)ΓC^{*}_{\text{max}}(M)^{\Gamma}, such that:

  1. (i)

    π\pi restricts to a *-homomorphism Cb()C_{b}(\mathbb{R})\rightarrow\mathcal{M};

  2. (ii)

    If |f(t)||g(t)||f(t)|\leq|g(t)| for all tt\in\mathbb{R}, then dom(g(D))dom(f(D))\textnormal{dom}(g(D))\subseteq\textnormal{dom}(f(D));

  3. (iii)

    If (fn)n(f_{n})_{n\in\mathbb{N}} is a sequence in C()C(\mathbb{R}) for which there exists fC()f^{\prime}\in C(\mathbb{R}) such that |fn(t)||f(t)||f_{n}(t)|\leq|f^{\prime}(t)| for all tt\in\mathbb{R}, and if fnf_{n} converge to a limit function fC()f\in C(\mathbb{R}) uniformly on compact subsets of \mathbb{R}, then fn(D)xf(D)xf_{n}(D)x\to f(D)x for each xdom(f(D))x\in\textnormal{dom}(f(D));

  4. (iv)

    Id(D)=D\textnormal{Id}(D)=D.

4.2. The wave operator

We now discuss the relationship between the wave operator formed using the functional calculus from Theorem 4.2 and the classical wave operator on L2L^{2}. Both of these operators can be viewed as bounded multipliers of the maximal Roe algebra Cmax(M)ΓC^{*}_{\textnormal{max}}(M)^{\Gamma}, and we will see that:

Proposition 4.3.

For each tt\in\mathbb{R}, we have

eL2itD=eitD(Cmax(M)Γ).e^{itD}_{L^{2}}=e^{itD}\in\mathcal{M}(C^{*}_{\textnormal{max}}(M)^{\Gamma}).

Let us begin by explaining both sides of the equation, starting with the right-hand side. For each tt\in\mathbb{R}, the functional calculus from Theorem 4.2 allows one to form a bounded adjointable operator eitDe^{itD} on the Hilbert module Cmax(M)ΓC^{*}_{\textnormal{max}}(M)^{\Gamma}. The resulting group of operators {eitD}t\{e^{itD}\}_{t\in\mathbb{R}} is strongly continuous in the sense of Theorem 4.2 (iii) and uniquely solves the wave equation on Cmax(M)ΓC^{*}_{\text{max}}(M)^{\Gamma}:

Lemma 4.4.

For any κ𝒮Γ\kappa\in\mathcal{S}^{\Gamma}, u(t)=eitDκu(t)=e^{itD}\kappa is the unique solution of the problem

dudt=iDu,u(0)=κ,\frac{du}{dt}=iDu,\quad u(0)=\kappa, (4.3)

with u:Cmax(M)Γu\colon\mathbb{R}\to C^{*}_{\text{max}}(M)^{\Gamma} a differentiable map taking values in dom(D)\textnormal{dom}(D).

Proof.

For each tt\in\mathbb{R}, the function seitss\mapsto e^{its} is a unitary in Cb()C_{b}(\mathbb{R}). Hence eitDe^{itD} is bounded adjointable and unitary. Let hnh_{n} be a sequence of positive real numbers converging to 0 as nn\rightarrow\infty. Then for each tt\in\mathbb{R}, the sequence of functions

fn(s)ei(t+hn)seitshnf_{n}(s)\coloneqq\frac{e^{i(t+h_{n})s}-e^{its}}{h_{n}}

converges to f(s)iseitsf(s)\coloneqq ise^{its} uniformly on compact subsets of \mathbb{R} in the limit nn\to\infty. Also, each fnf_{n} is bounded above by |1+s||1+s|. By Theorem 4.2 (iii), this implies (4.3).

For the uniqueness claim, let vv be another solution of (4.3) with v(0)=κv(0)=\kappa. For any fixed ss\in\mathbb{R} and 0ts0\leq t\leq s, set w(t)=eitDv(st)w(t)=e^{itD}v(s-t). Then we have

dwdt=iDeitDv(st)ieitDDv(st)=0.\frac{dw}{dt}=iDe^{itD}v(s-t)-ie^{itD}Dv(s-t)=0.

It follows that w(t)w(t) is constant for all tt, hence

v(s)=w(0)=w(s)=eisDκ=u(s).\displaystyle v(s)=w(0)=w(s)=e^{isD}\kappa=u(s).

On the other hand, we may apply the functional calculus on L2(E)L^{2}(E) to the essentially self-adjoint operator

D:L2(E)L2(E)D\colon L^{2}(E)\to L^{2}(E) (4.4)

to form the classical wave operator eL2itDe^{itD}_{L^{2}}, not to be confused with with the bounded adjointable operator eitDe^{itD}. The resulting operator group {eL2itD}t\{e^{itD}_{L^{2}}\}_{t\in\mathbb{R}} is strongly continuous in (L2(E))\mathcal{B}(L^{2}(E)) and uniquely solves the wave equation on L2(E)L^{2}(E): for every vCc(E)v\in C_{c}^{\infty}(E), we have

ddt(eL2itD(v))=iDeL2itD(v).\frac{d}{dt}(e^{itD}_{L^{2}}(v))=iDe^{itD}_{L^{2}}(v).

Via composition, eL2itDe^{itD}_{L^{2}} defines a map

eL2itD:𝒮Γ𝒮Γ.e^{itD}_{L^{2}}\colon\mathcal{S}^{\Gamma}\to\mathcal{S}^{\Gamma}. (4.5)

A standard argument involving the Sobolev embedding theorem now shows that for any Tκ𝒮ΓT_{\kappa}\in\mathcal{S}^{\Gamma}, the kernel of eL2itDTκe^{itD}_{L^{2}}\circ T_{\kappa} is smooth, in addition to having finite propagation and being Γ\Gamma-equivariant. Indeed, for any mm\in\mathbb{N}, the operator DmTκD^{m}\circ T_{\kappa} is bounded on L2(E)L^{2}(E), as M/ΓM/\Gamma is compact. Together with the fact that

DmeL2itDTκ=eL2itDDmTκ,D^{m}\circ e^{itD}_{L^{2}}\circ T_{\kappa}=e^{itD}_{L^{2}}\circ D^{m}\circ T_{\kappa},

Sobolev theory implies that the image of eL2itDTκe^{itD}_{L^{2}}\circ T_{\kappa} lies in the smooth sections, so that this operator also has smooth kernel. By Remark 2.9, (4.5) extends uniquely to a bounded multiplier of Cmax(M)ΓC^{*}_{\textnormal{max}}(M)^{\Gamma},

eL2itD:Cmax(M)ΓCmax(M)Γ.e^{itD}_{L^{2}}\colon C^{*}_{\textnormal{max}}(M)^{\Gamma}\to C^{*}_{\textnormal{max}}(M)^{\Gamma}. (4.6)

This is the operator on the left-hand side of Proposition 4.3.

Observe that if kk is a smooth, compactly supported Schwartz kernel on M×MM\times M, then we have the pointwise equality

(ddteL2itDk)(x,y)=(iDeL2itDk)(x,y),\Big{(}\frac{d}{dt}e^{itD}_{L^{2}}k\Big{)}(x,y)=(iDe^{itD}_{L^{2}}k)(x,y), (4.7)

where eL2itDke^{itD}_{L^{2}}k denotes the smooth Schwartz kernel of the composition eL2itDTke^{itD}_{L^{2}}\circ T_{k}. The proof of Proposition 4.3, involves a more general form of this observation for kernels in 𝒮Γ\mathcal{S}^{\Gamma}:

Lemma 4.5.

Let κ𝒮Γ\kappa\in\mathcal{S}^{\Gamma}, and let eL2itDκe^{itD}_{L^{2}}\kappa denote the smooth Schwartz kernel of eL2itDTκe^{itD}_{L^{2}}\circ T_{\kappa}. Then for every tt\in\mathbb{R} and x,yMx,y\in M, we have the pointwise equality

(ddteL2itDκ)(x,y)=(iDeL2itDκ)(x,y).\Big{(}\frac{d}{dt}e^{itD}_{L^{2}}\kappa\Big{)}(x,y)=(iDe^{itD}_{L^{2}}\kappa)(x,y). (4.8)

Moreover, the path teL2itDκt\mapsto e^{itD}_{L^{2}}\kappa is continuous with respect to the operator norm, and

limh0eL2i(t+h)DTκeL2itDTκhiDeL2itDTκ(L2(E))=0.\lim_{h\to 0}\,\,\,\bigg{\|}\frac{e^{i(t+h)D}_{L^{2}}\circ T_{\kappa}-e^{itD}_{L^{2}}\circ T_{\kappa}}{h}-iDe^{itD}_{L^{2}}\circ T_{\kappa}\bigg{\|}_{\mathcal{B}(L^{2}(E))}=0. (4.9)
Proof.

Fix tt\in\mathbb{R}, ϵ>0\epsilon>0, and xx, yMy\in M. Take ϕCc(M)\phi\in C_{c}^{\infty}(M) such that ϕ(z)=1\phi(z)=1 if zB2t(x)z\in B_{2t}(x), and ϕ(z)=0\phi(z)=0 if zM\B3t(x)z\in M\backslash B_{3t}(x). Since κ\kappa has finite propagation, the smooth kernel ϕκ\phi\kappa is compactly supported in M×MM\times M, so by (4.7) we have

(ddteL2itDϕκ)(x,y)=(iDeL2itDϕκ)(x,y).\Big{(}\frac{d}{dt}e^{itD}_{L^{2}}\phi\kappa\Big{)}(x,y)=(iDe^{itD}_{L^{2}}\phi\kappa)(x,y). (4.10)

The fact that eL2itDe^{itD}_{L^{2}} has propagation at most tt, together with the fact that

(1ϕ)κ(z,y)=0(1-\phi)\kappa(z,y)=0

for all zB2t(x)z\in B_{2t}(x), implies that for all wBt(x)w\in B_{t}(x) we have (eL2itD(1ϕ)κ)(w,y)=0(e^{itD}_{L^{2}}(1-\phi)\kappa)(w,y)=0. It follows that

(eL2itDκ)(w,y)=(eL2itDϕκ)(w,y).(e^{itD}_{L^{2}}\kappa)(w,y)=(e^{itD}_{L^{2}}\phi\kappa)(w,y). (4.11)

Taking w=xw=x and combining with (4.10) gives (4.8).

To establish norm continuity, it suffices to show that teL2itDκt\mapsto e^{itD}_{L^{2}}\kappa is norm-continuous at t=0t=0. For each xx, yy, and tt in an interval [t0,t0][-t_{0},t_{0}], we have

|eL2itDκ(x,y)κ(x,y)||t|sups[t0,t0]|iDeL2isDκ(x,y)|\Big{|}e^{itD}_{L^{2}}\kappa(x,y)-\kappa(x,y)\Big{|}\leq|t|\cdot\sup_{s\in[-t_{0},t_{0}]}\Big{|}iDe^{isD}_{L^{2}}\kappa(x,y)\Big{|} (4.12)

by the mean value theorem applied to (4.8). Since the operator iDeL2isDκiDe^{isD}_{L^{2}}\kappa is Γ\Gamma-equivariant with finite propagation, and MM is cocompact, there exists a compact subset KMK\subseteq M such that

supx,yM,s[t0,t0]|iDeL2isDκ(x,y)|=supx,yK,s[t0,t0]|iDeL2isDκ(x,y)|Ct0\sup_{\begin{subarray}{c}x,y\in M,\\ s\in[-t_{0},t_{0}]\end{subarray}}\Big{|}iDe^{isD}_{L^{2}}\kappa(x,y)\Big{|}=\sup_{\begin{subarray}{c}x,y\in K,\\ s\in[-t_{0},t_{0}]\end{subarray}}\Big{|}iDe^{isD}_{L^{2}}\kappa(x,y)\Big{|}\leq C_{t_{0}} (4.13)

for some constant Ct0C_{t_{0}}. Now since eL2itDκκe^{itD}_{L^{2}}\kappa-\kappa has finite propagation, its operator norm can be estimated using and (4.12) and (4.13):

eL2itDκκ(L2(E))Csupx,yM|eL2itDκ(x,y)κ(x,y)|CCt0|t|,\|e^{itD}_{L^{2}}\kappa-\kappa\|_{\mathcal{B}(L^{2}(E))}\leq C\cdot\sup_{x,y\in M}\Big{|}e^{itD}_{L^{2}}\kappa(x,y)-\kappa(x,y)\Big{|}\leq CC_{t_{0}}|t|,

for some constant CC depending on the propagation of κ\kappa. We obtain norm continuity by taking a limit t0t\to 0. The proof of (7.1) is a straightforward adaptation of this argument. ∎

With these preparations, we now prove Proposition 4.3.

Proof of Proposition 4.3.

Fix tt\in\mathbb{R}. Let Tκ𝒮ΓT_{\kappa}\in\mathcal{S}^{\Gamma} with smooth Schwartz kernel κ\kappa. We claim that the kernel eL2itDκe^{itD}_{L^{2}}\kappa satisfies the wave equation in Cmax(M)ΓC^{*}_{\textnormal{max}}(M)^{\Gamma}. To see this, first note that by equation (7.1) we have

limh0eL2i(t+h)DTκeL2itDTκhiDeL2itDTκ(L2(E))=0.\lim_{h\to 0}\,\,\,\bigg{\|}\frac{e^{i(t+h)D}_{L^{2}}\circ T_{\kappa}-e^{itD}_{L^{2}}\circ T_{\kappa}}{h}-iDe^{itD}_{L^{2}}\circ T_{\kappa}\bigg{\|}_{\mathcal{B}(L^{2}(E))}=0.

Now since the kernels eL2i(t+h)Dκe^{i(t+h)D}_{L^{2}}\kappa and DeL2itDκDe^{itD}_{L^{2}}\kappa each have propagation at most

rprop(κ)+|t+h|,r\coloneqq\textnormal{prop}(\kappa)+|t+h|,

by Remark 2.9 there exists a constant CrC_{r} such that the norm of

eL2i(t+h)DκeL2itDκhiDeL2itD(κ)\frac{e^{i(t+h)D}_{L^{2}}\kappa-e^{itD}_{L^{2}}\kappa}{h}-iDe^{itD}_{L^{2}}(\kappa)

in Cmax(M)ΓC^{*}_{\textnormal{max}}(M)^{\Gamma} is bounded above by CrC_{r} times its norm in (L2(E))\mathcal{B}(L^{2}(E)). Thus

limh0eL2i(t+h)DκeL2itDκhiDeL2itD(κ)max=0,\lim_{h\to 0}\,\,\,\bigg{\|}\frac{e^{i(t+h)D}_{L^{2}}\kappa-e^{itD}_{L^{2}}\kappa}{h}-iDe^{itD}_{L^{2}}(\kappa)\bigg{\|}_{\textnormal{max}}=0,

so the operator group {eL2itD}t\{e^{itD}_{L^{2}}\}_{t\in\mathbb{R}} solves the wave equation (4.3) in Cmax(M)ΓC^{*}_{\textnormal{max}}(M)^{\Gamma}. It follows from the uniqueness property in Lemma 4.4 that eL2itDe^{itD}_{L^{2}} and eitDe^{itD} coincide on 𝒮Γ\mathcal{S}^{\Gamma}, and hence are equal as elements of \mathcal{M}. ∎


5. Functoriality for the higher index

In this section, we discuss functoriality in the case of the maximal higher index from the point of view of folding maps and wave operators. This will prepare us for the proof of our main result, Theorem 1.1, in the next section.

Suppose M1M_{1} and M2M_{2} are two Galois covers of NN with deck transformation groups Γ1\Gamma_{1} and Γ2Γ1/H\Gamma_{2}\cong\Gamma_{1}/H, for some normal subgroup HH of Γ1\Gamma_{1}. Then the quotient homomorphism Γ1Γ2\Gamma_{1}\rightarrow\Gamma_{2} induces a natural surjective *-homomorphism between group algebras,

α:Γ1Γ2,i=1kcγiγii=1kcγi[γi],\alpha\colon\mathbb{C}\Gamma_{1}\rightarrow\mathbb{C}\Gamma_{2},\quad\sum_{i=1}^{k}c_{\gamma_{i}}\gamma_{i}\mapsto\sum_{i=1}^{k}c_{\gamma_{i}}[\gamma_{i}],

where [γ][\gamma] is the class of γΓ1\gamma\in\Gamma_{1} in Γ1/H\Gamma_{1}/H. In this setting, we have the following functoriality result for the maximal higher index, which follows from a theorem of Valette [17, Theorem 1.1]:

Theorem 5.1.

Let NN be a closed Riemannian manifold. Let DND_{N} be a first-order, self-adjoint elliptic differential operator acting on a bundle ENNE_{N}\rightarrow N. Let M1M_{1} and M2M_{2} be Galois covers of NN with deck transformation groups Γ1\Gamma_{1} and Γ2Γ1/H\Gamma_{2}\cong\Gamma_{1}/H respectively, for a normal subgroup HH of Γ1\Gamma_{1}. Let D1D_{1} and D2D_{2} be the lifts of DND_{N} to M1M_{1} and M2M_{2}. Then the map on KK-theory induced by α\alpha relates the maximal higher indices of D1D_{1} and D2D_{2}:

α(IndΓ1,maxD1)=IndΓ2,maxD2K(Cmax(Γ2)).\alpha_{*}(\operatorname{Ind}_{\Gamma_{1},\textnormal{max}}D_{1})=\operatorname{Ind}_{\Gamma_{2},\textnormal{max}}D_{2}\in K_{\bullet}\big{(}C^{*}_{\textnormal{max}}(\Gamma_{2})\big{)}.

This result was proved originally using KKKK-theory. To prepare for the proof of Theorem 1.1, we now give a proof of Theorem 5.1 using local properties of the wave operator and the folding map Ψ\Psi from the previous sections.

5.1. Higher index

We begin by recalling the definition of the maximal higher index of an equivariant elliptic operator. Let Γ\Gamma, MM, and DD be as in section 4.

Let 𝒬/Cmax(M)Γ.\mathcal{Q}\coloneqq\mathcal{M}/C^{*}_{\textnormal{max}}(M)^{\Gamma}. Consider the short exact sequence of CC^{*}-algebras

0Cmax(M)Γ𝒬0,0\rightarrow C^{*}_{\textnormal{max}}(M)^{\Gamma}\rightarrow\mathcal{M}\rightarrow\mathcal{Q}\rightarrow 0,

where \mathcal{M} is shorthand for the multiplier algebra (Cmax(M)Γ)\mathcal{M}\big{(}C^{*}_{\textnormal{max}}(M)^{\Gamma}\big{)}. This induces the following six-term exact sequence in KK-theory:

K0(Cmax(M)Γ){K_{0}(C^{*}_{\textnormal{max}}(M)^{\Gamma})}K0(){K_{0}(\mathcal{M})}K0(𝒬){K_{0}(\mathcal{Q})}K1(𝒬){K_{1}(\mathcal{Q})}K1(){K_{1}(\mathcal{M})}K1(Cmax(M)Γ),{K_{1}(C^{*}_{\textnormal{max}}(M)^{\Gamma}),}1\scriptstyle{\partial_{1}}0\scriptstyle{\partial_{0}}

where the connecting maps 0\partial_{0} and 1\partial_{1}, known as index maps, are defined as follows.

Definition 5.2.
  1. (i)

    0\partial_{0}: let uu be an invertible matrix over 𝒬\mathcal{Q} representing a class in K1(𝒬)K_{1}(\mathcal{Q}). Let vv be the inverse of uu. Let UU and VV be lifts of uu and vv to a matrix algebra over \mathcal{M}. Then the matrix

    W=(10U1)(1V01)(10U1)W=\begin{pmatrix}1&0\\ U&1\end{pmatrix}\begin{pmatrix}1&-V\\ 0&1\end{pmatrix}\begin{pmatrix}1&0\\ U&1\end{pmatrix}

    is invertible, and P=W(1000)W1P=W\begin{pmatrix}1&0\\ 0&0\end{pmatrix}W^{-1} is an idempotent. We define

    0[u][P][0001]K0(Cmax(M)Γ).\partial_{0}[u]\coloneqq\left[P\right]-\begin{bmatrix}0&0\\ 0&1\end{bmatrix}\in K_{0}\big{(}C^{*}_{\textnormal{max}}(M)^{\Gamma}\big{)}. (5.1)
  2. (ii)

    1\partial_{1}: let qq be an idempotent matrix over 𝒬\mathcal{Q} representing a class in K0(𝒬)K_{0}(\mathcal{Q}). Let QQ be a lift of qq to a matrix algebra over \mathcal{M}. Then e2πiQe^{2\pi iQ} is a unitary in the unitized algebra, and we define

    1[q][e2πiQ]K1(Cmax(M)Γ).\partial_{1}[q]\coloneqq\left[e^{2\pi iQ}\right]\in K_{1}(C^{*}_{\textnormal{max}}(M)^{\Gamma}). (5.2)

This construction is applied to the operator DD via the functional calculus from Theorem 4.2, as follows. Let χ:\chi\colon\mathbb{R}\rightarrow\mathbb{R} be a continuous, odd function such that

limx+χ(x)=1,\lim_{x\rightarrow+\infty}\chi(x)=1,

known as a normalizing function. Using Theorem 4.2, we obtain an element χ(D)\chi(D) in \mathcal{M}. We now have:

Lemma 5.3.

The class of χ(D)\chi(D) in /Cmax(M)Γ\mathcal{M}/C^{*}_{\textnormal{max}}(M)^{\Gamma} is invertible and independent of the choice of normalizing function χ\chi.

Proof.

Let 𝒮()\mathcal{S}(\mathbb{R}) denote the Schwartz space of functions \mathbb{R}\rightarrow\mathbb{C}. Then for every f𝒮()f\in\mathcal{S}(\mathbb{R}) with compactly supported Fourier transform f^\widehat{f}, the operator f(D)f(D) is given by a smooth kernel [18, Proposition 2.10]. Since every fC0()f\in C_{0}(\mathbb{R}) function is a uniform limit of such functions, the first part of Theorem 4.2 implies that for every such ff we have f(D)Cmax(M)Γf(D)\in C^{*}_{\textnormal{max}}(M)^{\Gamma}.

Now if χ\chi is a normalizing function, then χ21C0()\chi^{2}-1\in C_{0}(\mathbb{R}). Hence the class of χ(D)\chi(D) in /Cmax(M)Γ\mathcal{M}/C^{*}_{\textnormal{max}}(M)^{\Gamma} is invertible. Since any two normalizing functions differ by an element of C0()C_{0}(\mathbb{R}), this class is independent of the choice of χ\chi. ∎

Using this lemma, one computes that

χ(D)+12\frac{\chi(D)+1}{2}

is an idempotent modulo Cmax(M)ΓC^{*}_{\textnormal{max}}(M)^{\Gamma} and so defines element of K0(/Cmax(M)Γ)K_{0}\big{(}\mathcal{M}/C^{*}_{\textnormal{max}}(M)^{\Gamma}\big{)}. This leads us to the definition of the maximal higher index of DD:

Definition 5.4.

For i=1,2i=1,2, let i\partial_{i} be the connecting maps from Definition 5.2. The maximal higher index of DD is the element

1[χ(D)]K0(Cmax(M)Γ),\displaystyle\partial_{1}\left[\chi(D)\right]\in K_{0}\big{(}C^{*}_{\textnormal{max}}(M)^{\Gamma}\big{)},\quad if dimM is even,\displaystyle\textnormal{ if $\dim M$ is even},
0[χ(D)+12]K1(Cmax(M)Γ),\displaystyle\partial_{0}\left[\tfrac{\chi(D)+1}{2}\right]\in K_{1}\big{(}C^{*}_{\textnormal{max}}(M)^{\Gamma}\big{)},\quad if dimM is odd.\displaystyle\textnormal{ if $\dim M$ is odd}.

5.2. Functoriality

In this subsection, we return to the geometric setup described in subsection 2.3 and give a new proof of Theorem 5.1.

A key idea is to use the local nature of the wave operator to prove:

Proposition 5.5.

For all tt\in\mathbb{R}, we have Ψ(eitD1)=eitD2\Psi(e^{itD_{1}})=e^{itD_{2}}.

Proof.

For j=1,2j=1,2, let eL2itDje^{itD_{j}}_{L^{2}} be the wave operator on L2(Ej)L^{2}(E_{j}). By (4.6), this operator extends uniquely to a bounded multiplier of Cmax(Mj)ΓjC^{*}_{\textnormal{max}}(M_{j})^{\Gamma_{j}}. By Proposition 4.3, we have eL2itDj=eitDjje^{itD_{j}}_{L^{2}}=e^{itD_{j}}\in\mathcal{M}_{j}. Thus to prove this proposition, it suffices to show that Ψ(eL2itD1)=eL2itD2\Psi(e^{itD_{1}}_{L^{2}})=e^{itD_{2}}_{L^{2}}, as elements of Bfp(L2(Ej))ΓjB_{\textnormal{fp}}(L^{2}(E_{j}))^{\Gamma_{j}}.

As a notational convenience, we will write eitDje^{itD_{j}} for eL2itDjBfp(L2(Ej))Γje^{itD_{j}}_{L^{2}}\in B_{\textnormal{fp}}(L^{2}(E_{j}))^{\Gamma_{j}}. Let the open covers 𝒰M1\mathcal{U}_{M_{1}} and 𝒰M2\mathcal{U}_{M_{2}} be as in (3.3) and (3.4), so that by definition, there exists some ϵ>0\epsilon>0 such that each ball of diameter ϵ\epsilon in M2M_{2} is evenly covered with respect to π:M1M2\pi\colon M_{1}\to M_{2}.

Let {Vk}\{V_{k}\} be another open cover of M2M_{2} such that each VkV_{k} has diameter at most ϵ2\frac{\epsilon}{2} and such that any compact subset of M2M_{2} intersects only finitely many of the VkV_{k}. Let {ρk}\{\rho_{k}\} be a partition of unity subordinate to {Vk}\{V_{k}\}.

Choose a positive integer nn such that tn<ϵ8\frac{t}{n}<\frac{\epsilon}{8}. Now since

eitD1=(eitnD1)n,e^{itD_{1}}=\big{(}e^{i\frac{t}{n}D_{1}}\big{)}^{n},

and Ψ\Psi is a *-homomorphism, it suffices to show that Ψ(eitnD1)=eitnD2\Psi(e^{i\frac{t}{n}D_{1}})=e^{i\frac{t}{n}D_{2}}. Noting that any section uCc(E2)u\in C_{c}(E_{2}) can be written as a finite sum u=kρku,u=\sum_{k}\rho_{k}u, we have

Ψ(eitnD1)u=kΨ(eitnD1)(ρku).\Psi(e^{i\frac{t}{n}D_{1}})u=\sum_{k}\Psi(e^{i\frac{t}{n}D_{1}})(\rho_{k}u).

We first claim that for each kk we have

Ψ(eitnD1)(ρku)=eitnD2(ρku).\Psi(e^{i\frac{t}{n}D_{1}})(\rho_{k}u)=e^{i\frac{t}{n}D_{2}}(\rho_{k}u). (5.3)

To see this, note that the ball of radius ϵ8\frac{\epsilon}{8} around supp(ρku)\operatorname{supp}(\rho_{k}u) has diameter at most ϵ\epsilon and so is evenly covered with respect to π\pi. Since the definition of Ψ\Psi is independent of the choice of compatible partitions of unity by Proposition 3.3, we may work with a partition of unity {ϕj[g]}\{\phi_{j}^{[g]}\} for M2M_{2} subordinate to a cover 𝒰M2\mathcal{U}_{M_{2}}, such that Bϵ4(supp(ρku))Uj0B_{\frac{\epsilon}{4}}(\operatorname{supp}(\rho_{k}u))\subseteq U_{j_{0}} for some open set Uj0[s0]𝒰M2U_{j_{0}}^{[s_{0}]}\in\mathcal{U}_{M_{2}} and ϕj0[s0]1\phi_{j_{0}}^{[s_{0}]}\equiv 1 on Bϵ8(supp(ρku))B_{\frac{\epsilon}{8}}(\operatorname{supp}(\rho_{k}u)), for some s0Ss_{0}\in S and j0Ij_{0}\in I. By the definition of the folding map (3.6) applied to this choice of open cover, we have

Ψ(eitnD1)(ρku)\displaystyle\Psi(e^{i\frac{t}{n}D_{1}})(\rho_{k}u) =gΓ1,sSi,jIϕi[g](eitnD1ϕjs(ρku))\displaystyle=\sum_{\begin{subarray}{c}g\in\Gamma_{1},\\ s\in S\end{subarray}}\sum_{i,j\in I}\phi_{i}^{[g]}\big{(}e^{i\frac{t}{n}D_{1}}\phi_{j}^{s}(\rho_{k}u)\big{)}
=ϕj0[s0](eitnD1(π|Uj0s0(ρku)))\displaystyle=\phi_{j_{0}}^{[s_{0}]}\big{(}e^{i\frac{t}{n}D_{1}}\big{(}\pi|_{U_{j_{0}}^{s_{0}}}^{*}(\rho_{k}u)\big{)}\big{)}
=π(eitnD1(π|Uj0s0(ρku))).\displaystyle=\pi_{*}\big{(}e^{i\frac{t}{n}D_{1}}\big{(}\pi|_{U_{j_{0}}^{s_{0}}}^{*}(\rho_{k}u)\big{)}\big{)}.

Now the wave equation on M1M_{1} reads

t(eitnD1(π|Uj0s0(ρku)))=iD1(eitnD1(π|Uj0s0(ρku))).\frac{\partial}{\partial t}\big{(}e^{i\frac{t}{n}D_{1}}\big{(}\pi|_{U_{j_{0}}^{s_{0}}}^{*}(\rho_{k}u)\big{)}\big{)}=iD_{1}\big{(}e^{i\frac{t}{n}D_{1}}\big{(}\pi|_{U_{j_{0}}^{s_{0}}}^{*}(\rho_{k}u)\big{)}\big{)}.

Applying π\pi_{*} to both sides of this equation and using that D1D_{1} is the lift of D2D_{2}, we obtain

t(Ψ(eitnD1)(ρku))=iD2(Ψ(eitnD1)(ρku)),\frac{\partial}{\partial t}\big{(}\Psi\big{(}e^{i\frac{t}{n}D_{1}}\big{)}(\rho_{k}u)\big{)}=iD_{2}\big{(}\Psi\big{(}e^{i\frac{t}{n}D_{1}}\big{)}(\rho_{k}u)\big{)},

whence (5.3) follows from uniqueness of the solution to the wave equation on M2M_{2}.

Taking a sum over kk now yields

Ψ(eitnD1)u=kΨ(eitnD1)(ρku)=keitnD2(ρku)=eitnD2u.\Psi(e^{i\frac{t}{n}D_{1}})u=\sum_{k}\Psi(e^{i\frac{t}{n}D_{1}})(\rho_{k}u)=\sum_{k}e^{i\frac{t}{n}D_{2}}(\rho_{k}u)=e^{i\frac{t}{n}D_{2}}u.

Thus Ψ(eitnD1)=eitnD2\Psi(e^{i\frac{t}{n}D_{1}})=e^{i\frac{t}{n}D_{2}}, as bounded operators on L2(E2)L^{2}(E_{2}). By our previous remarks, this means that Ψ(eitD1)=eitD2\Psi(e^{it{D_{1}}})=e^{itD_{2}}. ∎

Applying Fourier inversion together with Proposition 5.5 leads to:

Proposition 5.6.

For any fC0()f\in C_{0}(\mathbb{R}) we have

Ψ(f(D1))=f(D2)2.\Psi(f(D_{1}))=f(D_{2})\in\mathcal{M}_{2}.
Proof.

Suppose first that f𝒮()f\in\mathcal{S}(\mathbb{R}) with compactly supported Fourier transform. By the Fourier inversion formula, we have

f(Dj)=12πf^(t)eitDj𝑑t,f(D_{j})=\frac{1}{2\pi}\int_{\mathbb{R}}\widehat{f}(t)e^{itD_{j}}\,dt, (5.4)

where the integral converges strongly in j\mathcal{M}_{j}. Now for any κCmax(M2)Γ2\kappa\in C^{*}_{\textnormal{max}}(M_{2})^{\Gamma_{2}}, Proposition 3.5 implies that there exists κ0Cmax(M1)Γ1\kappa_{0}\in C^{*}_{\textnormal{max}}(M_{1})^{\Gamma_{1}} such that Ψ(κ0)=κ\Psi(\kappa_{0})=\kappa. Since Ψ\Psi is a *-homomorphism,

Ψ(f(D1))(κ)\displaystyle\Psi(f(D_{1}))(\kappa) =Ψ(f(D1)κ0)=12πf^(t)Ψ(eitD1κ0)𝑑t.\displaystyle=\Psi(f(D_{1})\kappa_{0})=\frac{1}{2\pi}\int_{\mathbb{R}}\widehat{f}(t)\Psi(e^{itD_{1}}\kappa_{0})dt.

By Proposition 5.5, this equals

12πf^(t)eitD2κ𝑑t=f(D2)κ.\displaystyle\frac{1}{2\pi}\int_{\mathbb{R}}\widehat{f}(t)e^{itD_{2}}\kappa\,dt=f(D_{2})\kappa.

This proves the claim for f𝒮()f\in\mathcal{S}(\mathbb{R}). The general claim now follows from density of 𝒮()\mathcal{S}(\mathbb{R}) in C0()C_{0}(\mathbb{R}).

Proof of Theorem 5.1.

The expressions (5.1) and (5.2) show that, for j=1,2j=1,2, the higher index of DjD_{j} is represented by a matrix Aj(χ)A_{j}(\chi) whose entries are operators formed using functional calculus of DjD_{j}. More precisely, if the initial operator DND_{N} on NN is ungraded, as is typically the case for MM odd-dimensional, then

Aj(χ)=eπi(χ+1)(Dj).A_{j}(\chi)=e^{\pi i(\chi+1)}(D_{j}). (5.5)

When DjD_{j} is odd-graded with respect to a 2\mathbb{Z}_{2}-grading on the bundle Ej=Ej+EjE_{j}=E_{j}^{+}\oplus E_{j}^{-}, as typically occurs when dimN\dim N is even, we have a direct sum decomposition χ(Dj)=χ(Dj)+χ(Dj)\chi(D_{j})=\chi(D_{j})^{+}\oplus\chi(D_{j})^{-}. In this case, the index element is represented explicitly by the matrix

Aj(χ)=((1χ(Dj)χ(Dj)+)2χ(Dj)(1χ(Dj)+χ(Dj))χ(Dj)+(2χ(Dj)χ(Dj)+)(1χ(Dj)χ(Dj)+)χ(Dj)+χ(Dj)(2χ(Dj)+χ(Dj))1)A_{j}(\chi)=\left(\begin{smallmatrix}(1-\chi(D_{j})^{-}\chi(D_{j})^{+})^{2}&\,\,\chi(D_{j})^{-}(1-\chi(D_{j})^{+}\chi(D_{j})^{-})\\[4.30554pt] \chi(D_{j})^{+}(2-\chi(D_{j})^{-}\chi(D_{j})^{+})(1-\chi(D_{j})^{-}\chi(D_{j})^{+})&\,\,\chi(D_{j})^{+}\chi(D_{j})^{-}(2-\chi(D_{j})^{+}\chi(D_{j})^{-})-1\end{smallmatrix}\right) (5.6)

Observe that in either case, each entry of AjA_{j} is an operator of the form f(Dj)f(D_{j}), for some fC0()f\in C_{0}(\mathbb{R}) (modulo grading and the identity operator). By Proposition 5.6, Ψ\Psi maps each entry of A1A_{1} to the corresponding entry of A2A_{2}. Hence Ψ[A1]=[A2]\Psi_{*}[A_{1}]=[A_{2}], which proves the claim. ∎

Remark 5.7.

When Γ1=π1N\Gamma_{1}=\pi_{1}N and Γ2\Gamma_{2} is the trivial group, Theorem 5.1 reduces to the maximal version of Atiyah’s L2L^{2}-index theorem mentioned in section 1.

Atiyah’s original L2L^{2}-index theorem, which uses the von Neumann trace τ\tau instead of the folding map, can be proved by an argument along lines similar to the proof of Theorem 5.1.


6. Functoriality for the higher rho invariant

The higher index is a primary obstruction to the existence of positive scalar curvature metrics on a manifold. When the manifold is spin with positive scalar curvature, so that the higher index of the Dirac operator vanishes, one can define a secondary invariant called the higher rho invariant, introduced in [19, 12]. This is an obstruction to the inverse of the Dirac operator being local [4]. In this section we show that the higher rho invariant behaves functorially under the map ΨL,0\Psi_{L,0} from Definition 3.8.

6.1. Higher rho invariant

Consider the geometric situation in subsection 2.3, with the additional condition that the Riemannian manifold NN is spin with positive scalar curvature. The operator DND_{N} is then the Dirac operator acting on the spinor bundle ENE_{N}.

As the definition of the higher rho invariant is the same for either M1M_{1} or M2M_{2}, we will simply write MM, Γ\Gamma to mean either M1M_{1}, Γ1\Gamma_{1} or M2M_{2}, Γ2\Gamma_{2}. Similarly, DD will refer to either of the lifted operators D1D_{1} or D2D_{2} acting on the equivariant spinor bundles E1E_{1} or E2E_{2} lifted from ENE_{N}.

Let κ\kappa be the scalar curvature function of the lifted metric on MM, which is uniformly positive. Let :C(E)C(TME)\nabla\colon C^{\infty}(E)\rightarrow C^{\infty}(T^{*}M\otimes E) be the connection on EE induced by the Levi-Civita connection on MM. Recall that by the Lichnerowicz formula,

D2=+κ4.D^{2}=\nabla^{*}\nabla+\frac{\kappa}{4}.

Since κ\kappa is uniformly positive, D2D^{2} is strictly positive as an unbounded operator on the Hilbert module Cmax(M)ΓC^{*}_{\textnormal{max}}(M)^{\Gamma}. Thus we may use the functional calculus from Theorem 4.2 to form the operator

F0(D)D|D|,F_{0}(D)\coloneqq\frac{D}{|D|},

an element of the multiplier algebra =(Cmax(M)Γ)\mathcal{M}=\mathcal{M}(C^{*}_{\textnormal{max}}(M)^{\Gamma}). Observe that F0(D)+12\frac{F_{0}(D)+1}{2} is a projection in \mathcal{M}.

Since DD is invertible, there exists ϵ>0\epsilon>0 such that the spectrum of DD is contained in \(ϵ,ϵ)\mathbb{R}\backslash(-\epsilon,\epsilon). Let {Ft}t+\{F_{t}\}_{t\in\mathbb{R}^{+}} be a set of normalizing functions satisfying the following conditions:

  • FtF_{t} has compactly supported distributional Fourier transform for each tt;

  • diam(suppF^t)0\textnormal{diam}(\textnormal{supp}\,\widehat{F}_{t})\to 0 as tt\to\infty;

  • Ftx|x|F_{t}\to\frac{x}{|x|} uniformly on \(ϵ,ϵ)\mathbb{R}\backslash(-\epsilon,\epsilon) in the limit t0t\to 0.

In the limit tt\to\infty, the propagations of Ft(D)F_{t}(D) tend to 0. By Theorem 4.2 (i), as t0t\to 0, the operators Ft(D)F_{t}(D) converge to F0(D)F_{0}(D) in the norm of \mathcal{M}.

Define a path 0(Cmax(M)Γ)+\mathbb{R}^{\geq 0}\rightarrow(C^{*}_{\textnormal{max}}(M)^{\Gamma})^{+} given by

RD:tA(Ft),R_{D}\colon t\mapsto A(F_{t}), (6.1)

where the matrix A(Ft)A(F_{t}) is defined by (5.5) or (5.6) depending upon the dimension of NN (the subscript jj is omitted). Noting that

RD(0)=A(F0)={(0 00  1) if dimN is even,       1 if dimN is odd,R_{D}(0)=A(F_{0})=\begin{cases}\left(\begin{smallmatrix}0\,&\,0\\[4.30554pt] 0\,&\,\,1\end{smallmatrix}\right)&\textnormal{ if $\dim N$ is even},\\[4.30554pt] \,\,\,\,\,\,\,1&\textnormal{ if $\dim N$ is odd},\end{cases}

one sees that RDR_{D} is a matrix with entries in (CL,0,max(M)Γ)+\big{(}C^{*}_{L,0,\textnormal{max}}(M)^{\Gamma}\big{)}^{+}.

Definition 6.1.

The higher rho invariant of DD on the Riemannian manifold MM is

ρmax(D)=[RD]K(CL,0,max(M)Γ),\rho_{\textnormal{max}}(D)=\left[R_{D}\right]\in K_{\bullet}(C^{*}_{L,0,\textnormal{max}}(M)^{\Gamma}),

where =dimM (mod 2)\bullet=\dim M\textnormal{ (mod }2).

6.2. Functoriality

We are now ready to complete the proof of our main result, Theorem 1.1, using the tools we developed in sections 3, 4, and 5.

Recall from Definition 3.8 that we have a folding map at level of obstructions algebras,

ΨL,0:CL,0,max(M1)Γ1CL,0,max(M2)Γ2.\Psi_{L,0}\colon C^{*}_{L,0,\textnormal{max}}(M_{1})^{\Gamma_{1}}\to C^{*}_{L,0,\textnormal{max}}(M_{2})^{\Gamma_{2}}.

This map is well-defined because the folding map Ψ\Psi at the level of maximal equivariant Roe algebras preserves small propagation of operators, by Proposition 3.4. The induced map on KK-theory,

(ΨL,0):K(CL,0,max(M1)Γ1)K(CL,0,max(M2)Γ2),(\Psi_{L,0})_{*}\colon K_{\bullet}\left(C^{*}_{L,0,\textnormal{max}}(M_{1})^{\Gamma_{1}}\right)\rightarrow K_{\bullet}\left(C^{*}_{L,0,\textnormal{max}}(M_{2})^{\Gamma_{2}}\right),

implements functoriality of the maximal higher rho invariant.

Proof of Theorem 1.1.

For j=1,2j=1,2, let the higher rho invariants of DjD_{j} be denoted by ρmax(Dj)\rho_{\textnormal{max}}(D_{j}), as in Definition 6.1. By (6.1), this class is represented the path

RDj:tAj(Ft),R_{D_{j}}\colon t\mapsto A_{j}(F_{t}),

where the matrix AjA_{j} is as in (5.5) and (5.6). By Definition 3.8, the map (ΨL)(\Psi_{L})_{*} takes the class

[RD1]K(CL,0,max(M1)Γ1)[R_{D_{1}}]\in K_{\bullet}\big{(}C^{*}_{L,0,\textnormal{max}}(M_{1})^{\Gamma_{1}}\big{)}

to the class of the composed path

ΨRD1:tΨ(A1(Ft))\Psi\circ R_{D_{1}}\colon t\mapsto\Psi(A_{1}(F_{t}))

in K(CL,0,max(M2)Γ2)K_{\bullet}\big{(}C^{*}_{L,0,\textnormal{max}}(M_{2})^{\Gamma_{2}}\big{)}. Since each entry of AjA_{j} is an operator of the form f(Dj)f(D_{j}), for some fC0()f\in C_{0}(\mathbb{R}) (up to grading and the identity operator), Proposition 5.6 implies that for each t0t\geq 0, we have

ΨRD1(t)=Ψ(A1(Ft))=A2(Ft)=RD2(t).\Psi\circ R_{D_{1}}(t)=\Psi(A_{1}(F_{t}))=A_{2}(F_{t})=R_{D_{2}}(t).

It follows that (ΨL,0)(ρmax(D1))=ρmax(D2)K(CL,0,max(M2)Γ2)(\Psi_{L,0})_{*}(\rho_{\textnormal{max}}(D_{1}))=\rho_{\textnormal{max}}(D_{2})\in K_{\bullet}(C^{*}_{L,0,\textnormal{max}}(M_{2})^{\Gamma_{2}}). ∎


7. Generalizations to the non-cocompact setting

The methods in this paper can be used to establish analogous results in more general geometric settings. In this final section, we give two such generalizations, both involving non-cocompact actions.

We will work with the non-cocompact analogue of the geometric setup in subsection 2.3, so that the manifold NN is no longer assumed to be compact. We will assume throughout this section that the operator DND_{N} has unit propagation speed. In place of the finite partition of unity (3.2) used to define the folding map Ψ\Psi, we take a locally finite partition of unity 𝒰N\mathcal{U}_{N} whose elements are evenly covered with respect to the projections p1:M1Np_{1}\colon M_{1}\to N and p2:M2Np_{2}\colon M_{2}\to N, and with the property that any compact subset of NN intersects only finitely many elements of 𝒰N\mathcal{U}_{N}. The equivariant partitions of unity 𝒰M1\mathcal{U}_{M_{1}} and 𝒰M2\mathcal{U}_{M_{2}} of M1M_{1} and M2M_{2} are defined in the same way according to (3.3) and (3.4).

The local nature of the wave operator eitDe^{itD} means that Proposition 5.5 generalizes naturally to this setting:

Proposition 7.1.

Let NN, M1M_{1}, and M2M_{2} be as in this section, with NN not necessarily compact. Then for all tt\in\mathbb{R}, we have Ψ(eitD1)=eitD2\Psi(e^{itD_{1}})=e^{itD_{2}}.

Proof.

We adapt the proof of Proposition 5.5, indicating only what needs to be changed. Let the open covers 𝒰M1\mathcal{U}_{M_{1}} and 𝒰M2\mathcal{U}_{M_{2}} be as above. The difference now is that since NN may be non-compact, we cannot assume the existence of a uniformly positive covering diameter ϵ\epsilon as in the proof of Proposition 5.5.

Instead, the key point is to observe that for any fixed tt\in\mathbb{R} and uCc(E2)u\in C_{c}(E_{2}), the section eitD2ue^{itD_{2}}u is supported within the compact subset Bt(suppu)B_{t}(\operatorname{supp}u). Thus we can find ϵ>0\epsilon>0 such that for all xBt(suppu)x\in B_{t}(\operatorname{supp}u), the ball Bϵ(x)B_{\epsilon}(x) is evenly covered with respect to the projection π:M1M2\pi\colon M_{1}\to M_{2}. From here, we proceed precisely as in the proof of Proposition 5.5, with M2M_{2} replaced by Bt(suppu)B_{t}(\operatorname{supp}u), to show that Ψ(eitD1)u=eitD2u\Psi(e^{itD_{1}})u=e^{itD_{2}}u. Since tt and uu are arbitrary, we conclude. ∎

7.1. Operators invertible at infinity

In this subsection, suppose that the operator DND_{N} is invertible at infinity, meaning that there exists some compact subset ZNNZ_{N}\subseteq N on whose complement we have DN2aD_{N}^{2}\geq a for some a>0a>0. An important special case is when NN is spin, DND_{N} is the Dirac operator, and the metric gNg_{N} has uniformly positive scalar curvature outside of ZNZ_{N}.

For j=1,2j=1,2, the lifted operator DjD_{j} then satisfies the analogous relation Dj2aD_{j}^{2}\geq a on the complement of a cocompact, Γj\Gamma_{j}-invariant subset ZjMjZ_{j}\subseteq M_{j}. We can define a version of the higher index of DjD_{j}, localized around ZjZ_{j}, as follows (see also [20] and [10, section 3]).

For each R>0R>0, let Cmax(BR(Zj))ΓjC^{*}_{\textnormal{max}}(B_{R}(Z_{j}))^{\Gamma_{j}} be the maximal equivariant Roe algebra of the RR-neighborhood of ZjZ_{j}. Since BR(Zj)B_{R}(Z_{j}) is cocompact, this algebra is isomorphic to Cmax(Γj)𝒦C^{*}_{\textnormal{max}}(\Gamma_{j})\otimes\mathcal{K} by Remark 2.9. One can then show that for any fCc(a,a)f\in C_{c}(-a,a), we have

f(Dj)limRCmax(BR(Zj))Γj,f(D_{j})\in\lim_{R\rightarrow\infty}C^{*}_{\textnormal{max}}(B_{R}(Z_{j}))^{\Gamma_{j}},

where limRCmax(BR(Zj))Γj\lim_{R\rightarrow\infty}C^{*}_{\textnormal{max}}(B_{R}(Z_{j}))^{\Gamma_{j}} is the direct limit of these CC^{*}-algebras. Indeed, this limit algebra is isomorphic to Cmax(Γ)𝒦C^{*}_{\textnormal{max}}(\Gamma)\otimes\mathcal{K}, where 𝒦\mathcal{K} denotes the compact operators on a (not necessarily compact) fundamental domain of the Γ\Gamma-action. The construction from subsection 5.1 then gives an index element

IndΓj,maxDjK(Cmax(Γj)).\operatorname{Ind}_{\Gamma_{j},\textnormal{max}}D_{j}\in K_{\bullet}(C^{*}_{\textnormal{max}}(\Gamma_{j})).

Similar to the cocompact case, we have the following version of Theorem 5.1 for operators that are invertible at infinity:

Theorem 7.2.

Let NN be a Riemannian manifold and DND_{N} a first-order, self-adjoint elliptic differential operator acting on a bundle ENNE_{N}\rightarrow N. Assume that DND_{N} has unit propagation speed. Let M1M_{1} and M2M_{2} be Galois covers of NN with deck transformation groups Γ1\Gamma_{1} and Γ2Γ1/H\Gamma_{2}\cong\Gamma_{1}/H respectively, for a normal subgroup HH of Γ1\Gamma_{1}. Let D1D_{1} and D2D_{2} be the lifts of DND_{N} to M1M_{1} and M2M_{2} respectively. Then the map on KK-theory induced by the folding map Ψ\Psi relates the maximal higher indices of D1D_{1} and D2D_{2}:

Ψ(IndΓ1,maxD1)=IndΓ2,maxD2K(Cmax(Γ2)).\Psi_{*}(\operatorname{Ind}_{\Gamma_{1},\textnormal{max}}D_{1})=\operatorname{Ind}_{\Gamma_{2},\textnormal{max}}D_{2}\in K_{\bullet}\big{(}C^{*}_{\textnormal{max}}(\Gamma_{2})\big{)}.
Proof.

Analogous to the proof of Theorem 5.1, with Proposition 5.5 replaced by Proposition 7.1. ∎

7.2. Manifolds with equivariantly bounded geometry

A second generalization involves non-cocompact coverings that satisfy certain additional geometric conditions. Under these conditions we obtain generalizations of both Theorems 5.1 and 1.1. In contrast to Theorem 7.2, here we do not require invertibility of the operator at infinity, and the Roe algebra we work with does not need to be defined in terms of cocompact sets as in subsection 7.1.

Again, suppose that we are in the setup of subsection 2.3, but with NN not necessarily compact. We impose two conditions on the geometry of NN and the Γi\Gamma_{i}-action on MiM_{i} (see also [11, subsection 2.1]):

  1. A.

    The Riemannian manifold NN has positive injectivity radius, and its curvature tensor is uniformly bounded across NN along with all of its derivatives.

  2. B.

    For j=1,2j=1,2, there exists a fundamental domain 𝒟j\mathcal{D}_{j} for the action of Γj\Gamma_{j} on MjM_{j} such that

    l(γ)d(𝒟j,γ𝒟j),l(\gamma)\rightarrow\infty\implies d(\mathcal{D}_{j},\gamma\mathcal{D}_{j})\to\infty,

    where l:Γjl\colon\Gamma_{j}\to\mathbb{N} is a fixed length function and dd is the Riemannian distance on MjM_{j}.

It follows from these two assumptions and [11, Proposition 2.14] that the maximal equivariant Roe algebra Cmax(Mj)ΓjC^{*}_{\textnormal{max}}(M_{j})^{\Gamma_{j}} is well-defined, along with a subalgebra Cmax,u(Mj)ΓjC^{*}_{\textnormal{max},u}(M_{j})^{\Gamma_{j}} called the maximal equivariant uniform Roe algebra. Using the latter, we constructed in [11] a version of the functional calculus suitable for the maximal setting.

We briefly recall the definition of Cmax,u(Mj)ΓjC^{*}_{\textnormal{max},u}(M_{j})^{\Gamma_{j}}. Let 𝒮uΓj\mathcal{S}_{u}^{\Gamma_{j}} be the *-subalgebra of 𝒮Γj\mathcal{S}^{\Gamma_{j}} (see section 4) whose kernels have uniformly bounded derivatives of all orders. In other words, an element of 𝒮uΓj\mathcal{S}_{u}^{\Gamma_{j}} is a bounded operator on L2(Ej)L^{2}(E_{j}) given by a Schwartz kernel κCb(EjEj)\kappa\in C_{b}^{\infty}(E_{j}\boxtimes E_{j}^{*}) such that

  1. (i)

    κ\kappa has finite propagation;

  2. (ii)

    κ(x,y)=κ(γx,γy)\kappa(x,y)=\kappa(\gamma x,\gamma y) for all γΓj\gamma\in\Gamma_{j};

  3. (iii)

    Each covariant derivative of κ\kappa is uniformly bounded over MjM_{j}.

Then Cmax,u(Mj)ΓjC^{*}_{\textnormal{max},u}(M_{j})^{\Gamma_{j}} is defined to be the closure of 𝒮uΓj\mathcal{S}_{u}^{\Gamma_{j}} in Cmax(Mj)ΓjC^{*}_{\textnormal{max}}(M_{j})^{\Gamma_{j}}.

Similar to section 4, Cmax,u(Mj)ΓjC^{*}_{\textnormal{max},u}(M_{j})^{\Gamma_{j}} can be viewed as right Hilbert module over itself, with inner product and multiplication being defined by the same formula (4.1). The operator DjD_{j} can be viewed as a densely defined symmetric operator on Cmax,u(Mj)ΓjC^{*}_{\textnormal{max},u}(M_{j})^{\Gamma_{j}}. By [11, Theorem 3.1], this operator is regular and so admits a functional calculus. The construction from subsection 5.1 then allows one to define the maximal equivariant uniform index of DjD_{j},

indexΓj,max,uDjK(Cmax,u(Mj)Γj).\operatorname{index}_{\Gamma_{j},\textnormal{max},u}D_{j}\in K_{\bullet}(C^{*}_{\textnormal{max},u}(M_{j})^{\Gamma_{j}}).

The corresponding versions of the higher rho invariant can be defined as in subsection 6.1, and we denote this by

ρmax,u(Dj)K(CL,0,max,u(Mj)Γj),\rho_{\textnormal{max,u}}(D_{j})\in K_{\bullet}(C^{*}_{L,0,\textnormal{max,u}}(M_{j})^{\Gamma_{j}}),

where CL,0,max,u(Mj)ΓjC^{*}_{L,0,\textnormal{max,u}}(M_{j})^{\Gamma_{j}} is defined analogously to CL,0,max(Mj)ΓjC^{*}_{L,0,\textnormal{max}}(M_{j})^{\Gamma_{j}}, with Cmax,u(Mj)ΓjC^{*}_{\textnormal{max},u}(M_{j})^{\Gamma_{j}} in place of Cmax(Mj)ΓjC^{*}_{\textnormal{max}}(M_{j})^{\Gamma_{j}}. Constructions analogous to those in subsection 3.3 give way to folding maps at KK-theory level, which we denote by

(Ψu):K(Cmax,u(M1)Γ1)K(Cmax,u(M2)Γ2),(\Psi_{u})_{*}\colon K_{\bullet}(C^{*}_{\textnormal{max},u}(M_{1})^{\Gamma_{1}})\rightarrow K_{\bullet}(C^{*}_{\textnormal{max},u}(M_{2})^{\Gamma_{2}}),
(ΨL,0,u):K(CL,0,max,u(M1)Γ1)K(CL,0,max,u(M2)Γ2).(\Psi_{L,0,u})_{*}\colon K_{\bullet}(C^{*}_{L,0,\textnormal{max},u}(M_{1})^{\Gamma_{1}})\rightarrow K_{\bullet}(C^{*}_{L,0,\textnormal{max},u}(M_{2})^{\Gamma_{2}}).

Further, Proposition 4.3 generalizes naturally to manifolds satisfying conditions A and B above:

Proposition 7.3.

For each tt\in\mathbb{R}, we have

eL2itDj=eitDjj(Cmax,u(Mj)Γj),e^{itD_{j}}_{L^{2}}=e^{itD_{j}}\in\mathcal{M}_{j}(C^{*}_{\textnormal{max},u}(M_{j})^{\Gamma_{j}}),

with notation as in subsection 4.2.

Proof.

The proof proceeds exactly as for Proposition 4.3, but with Lemma 4.5 replaced by Lemma 7.4 below. ∎

Lemma 7.4.

Let NN, M1M_{1}, and M2M_{2} be as in this section, with NN not necessarily compact. Let κ𝒮uΓj\kappa\in\mathcal{S}_{u}^{\Gamma_{j}}, and let eL2itDjκe^{itD_{j}}_{L^{2}}\kappa denote the smooth Schwartz kernel of eL2itDjTκe^{itD_{j}}_{L^{2}}\circ T_{\kappa}. Then the path teL2itDjκt\mapsto e^{itD_{j}}_{L^{2}}\kappa is continuous with respect to the operator norm, and we have

limh0eL2i(t+h)DjTκeL2itDjTκhiDjeL2itDjTκ(L2(E))=0.\lim_{h\to 0}\,\,\,\bigg{\|}\frac{e^{i(t+h)D_{j}}_{L^{2}}\circ T_{\kappa}-e^{itD_{j}}_{L^{2}}\circ T_{\kappa}}{h}-iD_{j}e^{itD_{j}}_{L^{2}}\circ T_{\kappa}\bigg{\|}_{\mathcal{B}(L^{2}(E))}=0. (7.1)
Proof.

Similar to equation (4.12), we have, for each x,yMjx,y\in M_{j} and t[t0,t0]t\in[-t_{0},t_{0}],

|eL2itDjκ(x,y)κ(x,y)||t|sups[t0,t0]|iDjeL2isDjκ(x,y)|.\Big{|}e^{itD_{j}}_{L^{2}}\kappa(x,y)-\kappa(x,y)\Big{|}\leq|t|\cdot\sup_{s\in[-t_{0},t_{0}]}\Big{|}iD_{j}e^{isD_{j}}_{L^{2}}\kappa(x,y)\Big{|}\,.

Since MjM_{j} has bounded Riemannian geometry, it follows from Sobolev theory that the operator iDjeL2isDjκ(x,y)iD_{j}e^{isD_{j}}_{L^{2}}\kappa(x,y) has uniformly bounded smooth kernel, along with all derivatives. Indeed, since eL2isDje^{isD_{j}}_{L^{2}} is unitary, |iDjeL2isDjκ(x,y)|\big{|}iD_{j}e^{isD_{j}}_{L^{2}}\kappa(x,y)\big{|} is bounded above by a constant C1C_{1} independent of ss, xx, and yy. It follows that we can estimate the L2L^{2}-norm of the finite-propagation operator eL2itDjκκe^{itD_{j}}_{L^{2}}\kappa-\kappa by

eL2itDjκκ(L2(E))C2supx,yM|eL2itDjκ(x,y)κ(x,y)|C1C2,\|e^{itD_{j}}_{L^{2}}\kappa-\kappa\|_{\mathcal{B}(L^{2}(E))}\leq C_{2}\cdot\sup_{x,y\in M}\Big{|}e^{itD_{j}}_{L^{2}}\kappa(x,y)-\kappa(x,y)\Big{|}\leq C_{1}C_{2},

for some other constant C2C_{2}. We can then finish the proof as in Lemma 4.5. ∎

With this in hand, we arrive (via an obvious analogue of Proposition 5.6) at the following generalizations of Theorems 5.1 and 1.1.

Theorem 7.5.

Let NN be a Riemannian manifold and M1M_{1} and M2M_{2} be Galois covers of NN with deck transformation groups Γ1\Gamma_{1} and Γ2Γ1/H\Gamma_{2}\cong\Gamma_{1}/H respectively, for a normal subgroup HH of Γ1\Gamma_{1}. Suppose that the conditions A and B in this subsection are satisfied. Let DND_{N} be a first-order, self-adjoint elliptic differential operator acting on a bundle ENNE_{N}\rightarrow N, and assume that DND_{N} has unit propagation speed. Let D1D_{1} and D2D_{2} be the lifts of DND_{N} to M1M_{1} and M2M_{2} respectively. We then have:

(Ψu,)(indexΓ1,max,uD1)\displaystyle(\Psi_{u,*})(\operatorname{index}_{\Gamma_{1},\textnormal{max,u}}D_{1}) =indexΓ2,max,uD2K(Cmax,u(M2)Γ2),\displaystyle=\operatorname{index}_{\Gamma_{2},\textnormal{max,u}}D_{2}\in K_{\bullet}\big{(}C^{*}_{\textnormal{max},u}(M_{2})^{\Gamma_{2}}\big{)},
(Ψu,)(indexΓ1,maxD1)\displaystyle(\Psi_{u,*})(\operatorname{index}_{\Gamma_{1},\textnormal{max}}D_{1}) =indexΓ2,maxD2K(Cmax(M2)Γ2).\displaystyle=\operatorname{index}_{\Gamma_{2},\textnormal{max}}D_{2}\in K_{\bullet}\big{(}C^{*}_{\textnormal{max}}(M_{2})^{\Gamma_{2}}\big{)}.
Theorem 7.6.

Suppose (N,gN)(N,g_{N}) is a spin Riemannian manifold with uniformly positive scalar curvature. Let M1M_{1} and M2M_{2} be Galois covers of MM with deck transformation groups Γ1\Gamma_{1} and Γ2Γ1/H\Gamma_{2}\cong\Gamma_{1}/H respectively, for a normal subgroup HH of Γ1\Gamma_{1}. Suppose that the conditions A and B in this subsection are satisfied. Let DND_{N} be the Dirac operator on NN (with unit propagation speed). Let D1D_{1} and D2D_{2} be the lifts of DND_{N} to M1M_{1} and M2M_{2} respectively. We then have:

(ΨL,0,u)(ρmax,u(D1))=ρmax,u(D2)K(CL,0,max,u(M2)Γ2),\displaystyle(\Psi_{L,0,u})_{*}\big{(}\rho_{\textnormal{max,u}}(D_{1})\big{)}=\rho_{\textnormal{max,u}}(D_{2})\in K_{\bullet}\big{(}C^{*}_{L,0,\textnormal{max,u}}(M_{2})^{\Gamma_{2}}\big{)},
(ΨL,0,u)(ρmax(D1))=ρmax(D2)K(CL,0,max(M2)Γ2).\displaystyle(\Psi_{L,0,u})_{*}\big{(}\rho_{\textnormal{max}}(D_{1})\big{)}=\rho_{\textnormal{max}}(D_{2})\in K_{\bullet}\big{(}C^{*}_{L,0,\textnormal{max}}(M_{2})^{\Gamma_{2}}\big{)}.
Remark 7.7.

The second equality in Theorem 7.5 follows from the observation that the natural inclusion Cmax,u(Mj)ΓjCmax(Mj)ΓjC^{*}_{\textnormal{max},u}(M_{j})^{\Gamma_{j}}\hookrightarrow C^{*}_{\textnormal{max}}(M_{j})^{\Gamma_{j}} relates indexΓj,max,uDj\operatorname{index}_{\Gamma_{j},\textnormal{max,u}}D_{j} to indexΓj,maxDj\operatorname{index}_{\Gamma_{j},\textnormal{max}}D_{j}. The second equality in Theorem 7.6 follows from a similar relationship at the level of obstruction algebras.



References

  • [1] Paul Baum and Alain Connes. KK-theory for discrete groups. In Operator algebras and applications, Vol. 1, volume 135 of London Math. Soc. Lecture Note Ser., pages 1–20. Cambridge Univ. Press, Cambridge, 1988.
  • [2] Paul Baum, Alain Connes, and Nigel Higson. Classifying space for proper actions and KK-theory of group CC^{\ast}-algebras. In CC^{\ast}-algebras: 1943–1993 (San Antonio, TX, 1993), volume 167 of Contemp. Math., pages 240–291. Amer. Math. Soc., Providence, RI, 1994.
  • [3] Stanley Chang, Shmuel Weinberger, and Guoliang Yu. Positive scalar curvature and a new index theory for noncompact manifolds. J. Geom. Phys., 149:103575, 22, 2020.
  • [4] Xiaoman Chen, Hongzhi Liu, and Guoliang Yu. Higher ρ\rho invariant is an obstruction to the inverse being local. J. Geom. Phys., 150:103592, 18, 2020.
  • [5] Xiaoman Chen, Jinmin Wang, Zhizhang Xie, and Guoliang Yu. Delocalized eta invariants, cyclic cohomology and higher rho invariants. arXiv:1901.02378.
  • [6] Alain Connes. Noncommutative geometry. Academic Press, Inc., San Diego, CA, 1994.
  • [7] Kenneth R. Davidson. CC^{*}-algebras by example, volume 6 of Fields Institute Monographs. American Mathematical Society, Providence, RI, 1996.
  • [8] Guihua Gong, Qin Wang, and Guoliang Yu. Geometrization of the strong Novikov conjecture for residually finite groups. J. Reine Angew. Math., 621:159–189, 2008.
  • [9] Hao Guo, Peter Hochs, and Varghese Mathai. Coarse geometry and Callias quantisation. Trans. Amer. Math. Soc., to appear.
  • [10] Hao Guo, Peter Hochs, and Varghese Mathai. Positive scalar curvature and an equivariant Callias-type index theorem for proper actions. Ann. K-theory., to appear.
  • [11] Hao Guo, Zhizhang Xie, and Guoliang Yu. A Lichnerowicz vanishing theorem for the maximal Roe algebra. arXiv:1905.12299. Submitted.
  • [12] Nigel Higson and John Roe. Mapping surgery to analysis. III. Exact sequences. KK-Theory, 33(4):325–346, 2005.
  • [13] G. G. Kasparov. Equivariant KKKK-theory and the Novikov conjecture. Invent. Math., 91(1):147–201, 1988.
  • [14] Dan Kucerovsky. A short proof of an index theorem. Proc. Amer. Math. Soc., 129(12):3729–3736, 2001.
  • [15] E. C. Lance. Hilbert CC^{*}-modules, volume 210 of London Mathematical Society Lecture Note Series. Cambridge University Press, Cambridge, 1995. A toolkit for operator algebraists.
  • [16] Hongzhi Liu, Xiang Tang, Zhizhang Xie, Yi-Jun Yao, and Guoliang Yu. Higher rho invariants and rigidity theorems for relative eta invariants. Preprint.
  • [17] Guido Mislin and Alain Valette. Proper group actions and the Baum-Connes conjecture. Advanced Courses in Mathematics. CRM Barcelona. Birkhäuser Verlag, Basel, 2003.
  • [18] John Roe. An index theorem on open manifolds. I, II. J. Differential Geom., 27(1):87–113, 115–136, 1988.
  • [19] John Roe. Index theory, coarse geometry, and topology of manifolds, volume 90 of CBMS Regional Conference Series in Mathematics. Published for the Conference Board of the Mathematical Sciences, Washington, DC; by the American Mathematical Society, Providence, RI, 1996.
  • [20] John Roe. Positive curvature, partial vanishing theorems and coarse indices. Proc. Edinb. Math. Soc. (2), 59(1):223–233, 2016.
  • [21] Thomas Schick and Mehran Seyedhosseini. On an index theorem of Chang, Weinberger and Yu. Münst. J. Math., to appear.
  • [22] Yanli Song and Xiang Tang. Higher orbit integrals, cyclic cocyles, and KK-theory of reduced group CC^{*}-algebra. arXiv:1910.00175.
  • [23] Bai-Ling Wang and Hang Wang. Localized index and L2L^{2}-Lefschetz fixed-point formula for orbifolds. J. Differential Geom., 102(2):285–349, 2016.
  • [24] Jinmin Wang, Zhizhang Xie, and Guoliang Yu. Approximations of delocalized eta invariants by their finite analogues. arXiv:2003.03401.
  • [25] Shmuel Weinberger, Zhizhang Xie, and Guoliang Yu. Additivity of higher rho invariants and nonrigidity of topological manifolds. Comm. Pure Appl. Math., 74(1):3–113, 2021.
  • [26] Rufus Willett and Guoliang Yu. Higher Index Theory. Cambridge Studies in Advanced Mathematics. Cambridge University Press, 2020.
  • [27] Zhizhang Xie and Guoliang Yu. Higher rho invariants and the moduli space of positive scalar curvature metrics. Adv. Math., 307:1046–1069, 2017.
  • [28] Zhizhang Xie and Guoliang Yu. Delocalized Eta Invariants, Algebraicity, and K-Theory of Group C*-Algebras. Int. Math. Res. Not. IMRN, 08 2019. rnz170.
  • [29] Zhizhang Xie, Guoliang Yu, and Rudolf Zeidler. On the range of the relative higher index and the higher rho-invariant for positive scalar curvature. arXiv:1712.03722.
  • [30] Guoliang Yu. A characterization of the image of the Baum-Connes map. In Quanta of maths, volume 11 of Clay Math. Proc., pages 649–657. Amer. Math. Soc., Providence, RI, 2010.