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

The Mayer-Vietoris Sequence for the Analytic Structure Group

Paul Siegel
Abstract.

We develop a Mayer-Vietoris sequence for the analytic structure group explored by Higson and Roe. Using explicit formulas for Mayer-Vietoris boundary maps, we give a new proof and generalizations of Roe’s partitioned manifold index theorem ([8]).

1. Introduction

Let XX be a proper metric space. Higson, Roe, and others have studied a map

μ:Kp(X)Kp(C(X))\mu\colon K_{p}(X)\to K_{p}(C^{*}(X))

from the K-homology of XX to the K-theory of the coarse C*-algebra of XX called the coarse assembly map. It fits into a long exact sequence

(1.1) Kp+1(C(X))Sp(X)Kp(X)Kp(C(X))\to K_{p+1}(C^{*}(X))\to S_{p}(X)\to K_{p}(X)\to K_{p}(C^{*}(X))\to

with a group Sp(X)S_{p}(X) called the analytic structure group for its relationship with the structure set in algebraic topology (established in [4], [5], and [6]).

The goal of this note is to develop a Mayer-Vietoris sequence for the analytic structure group which is compatible with the ordinary Mayer-Vietoris sequence in K-homology and the Mayer-Vietoris sequence for coarse spaces developed in [7]. Thus for an appropriate decomposition X=Y1Y2X=Y_{1}\cup Y_{2}, these Mayer-Vietoris sequences fit together with (1.1) to form a braid diagram:


Kp(Y1)Kp(Y2)\textstyle{K_{p}(Y_{1})\oplus K_{p}(Y_{2})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Kp(C(Y1))Kp(C(Y2))\textstyle{K_{p}(C^{*}(Y_{1}))\oplus K_{p}(C^{*}(Y_{2}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Kp(C(X))\textstyle{K_{p}(C^{*}(X))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Sp1(X)\textstyle{S_{p-1}(X)}Kp(C(Y1Y2))\textstyle{K_{p}(C^{*}(Y_{1}\cap Y_{2}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Kp(X)\textstyle{K_{p}(X)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Sp1(Y1)Sp1(Y2)\textstyle{S_{p-1}(Y_{1})\oplus S_{p-1}(Y_{2})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Kp1(C(Y1Y2))\textstyle{K_{p-1}(C^{*}(Y_{1}\cap Y_{2}))}Sp(X)\textstyle{S_{p}(X)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Sp1(Y1Y2)\textstyle{S_{p-1}(Y_{1}\cap Y_{2})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Kp1(Y1Y2)\textstyle{K_{p-1}(Y_{1}\cap Y_{2})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Kp1(Y1)Kp1(Y2)\textstyle{K_{p-1}(Y_{1})\oplus K_{p-1}(Y_{2})}

Exploiting connections between the coarse assembly map and index theory for elliptic operators, we use this braid diagram to give a new proof of a theorem of Roe ([8]). Roe posed and solved an index problem on a non-compact manifold MM which is partitioned by a compact hypersurface NN, meaning MM is the union of two submanifolds whose common boundary is NN. Our approach yields new generalizations of Roe’s theorem to equivariant indices and to non-compact partitioning hypersurfaces, and as an application we give a simple proof of a classical theorem of Gromov and Lawson ([1]) about topological obstructions to the existence of positive scalar curvature metrics.

2. The Assembly Map and the Analytic Structure Group

We begin by reviewing the construction of the coarse assembly map and the analytic structure group following the textbook [3].

2.1. K-Homology

The source of the coarse assembly map is the K-homology of the space XX; we now review a C*-algebraic definition of K-homology.

Definition 2.1.

Let XX be a locally compact Hausdorff space, let HH be a Hilbert space, and let ρ:C0(X)𝔹(H)\rho\colon C_{0}(X)\to\mathbb{B}(H) be a C*-algebra representation.

  • The dual algebra of XX is the C*-algebra 𝔇ρ(X)\mathfrak{D}^{*}_{\rho}(X) consisting of those operators T𝔹(H)T\in\mathbb{B}(H) such that TT commutes with ρ(f)\rho(f) modulo compact operators for each fC0(X)f\in C_{0}(X). Such operators are said to be pseudolocal

  • The locally compact algebra is the C*-ideal ρ(X)\mathfrak{C}^{*}_{\rho}(X) in 𝔇(X)\mathfrak{D}^{*}(X) consisting of those operators T𝔹(H)T\in\mathbb{B}(H) such that ρ(f)T\rho(f)T and Tρ(f)T\rho(f) are compact for each fC0(X)f\in C_{0}(X). Such operators are said to be locally compact

  • Let 𝔔ρ(X)\mathfrak{Q}^{*}_{\rho}(X) denote the quotient C*-algebra 𝔇ρ(X)/ρ(X)\mathfrak{D}^{*}_{\rho}(X)/\mathfrak{C}^{*}_{\rho}(X).

We will often use the notation “\sim” to denote equality modulo compact operators; thus T𝔇ρ(X)T\in\mathfrak{D}^{*}_{\rho}(X) if and only if [T,ρ(f)]0[T,\rho(f)]\sim 0 for every fC0(X)f\in C_{0}(X) and Tρ(X)T\in\mathfrak{C}^{*}_{\rho}(X) if and only if Tρ(f)ρ(f)T0T\rho(f)\sim\rho(f)T\sim 0 for every fC0(X)f\in C_{0}(X).

Note that the dual algebra, the locally compact algebra, and their quotient each depend on the representation ρ\rho. However, their K-theory groups are independent of ρ\rho so long as C0(X)C_{0}(X) is separable (equivalently, XX is second countable), HH is separable, and ρ\rho is ample (meaning ρ\rho is nondegenerate and ρ(f)\rho(f) is a compact operator only if f=0f=0 for fC0(X)f\in C_{0}(X)). See chapter 5 of [3] for further details. Because of this we will often abuse notation and write 𝔇(X)\mathfrak{D}^{*}(X), (X)\mathfrak{C}^{*}(X) and 𝔔(X)\mathfrak{Q}^{*}(X) with the understanding that the representations used to define these C*-algebras are ample and the Hilbert spaces are separable.

Definition 2.2.

Let XX be a second countable locally compact Hausdorff space. The ppth K-homology group of XX, pp\in\mathbb{Z}, is defined to be:

Kp(X)=K1p(𝔔(X))K_{p}(X)=K_{1-p}(\mathfrak{Q}^{*}(X))
Remark 2.3.

In fact we shall see that Kp((X))=0K_{p}(\mathfrak{C}^{*}(X))=0, so we could have defined Kp(X)K_{p}(X) to be K1p(𝔇(X))K_{1-p}(\mathfrak{D}^{*}(X)) as is done in chapter 5 of [3].

The K-homology groups form a generalized homology theory (in the sense of the Eilenberg-Steenrod axioms) which is naturally dual to topological K-theory. In particular they enjoy a certain functoriality property which can be understood at the level of dual algebras as follows. To begin, let ϕ:𝒰Y\phi\colon\mathcal{U}\to Y be a continuous proper map from an open subset 𝒰\mathcal{U} of XX to YY. Extend ϕ\phi to a map ϕ+:X+Y+\phi^{+}\colon X^{+}\to Y^{+} between one-point compactifications by sending the complement of 𝒰\mathcal{U} in XX to the point at infinity in Y+Y^{+}. Then ϕ+\phi^{+} induces a *-homomorphism C(Y+)C(U+)C(X+)C(Y^{+})\to C(U^{+})\subseteq C(X^{+}) given by ffϕ+f\mapsto f\circ\phi^{+}, and we define ϕ\phi^{*} to be the restriction of this map to C0(Y)C_{0}(Y).

Definition 2.4.

Let XX and YY be second countable locally compact Hausdorff spaces, let ρX:C0(X)𝔹(HX)\rho_{X}\colon C_{0}(X)\to\mathbb{B}(H_{X}) and ρY:C0(Y)𝔹(HY)\rho_{Y}\colon C_{0}(Y)\to\mathbb{B}(H_{Y}) be ample representations on separable Hilbert spaces, and let ϕ:𝒰Y\phi\colon\mathcal{U}\to Y be a continuous proper map defined on an open subset 𝒰X\mathcal{U}\subseteq X. An isometry V:HXHYV\colon H_{X}\to H_{Y} topologically covers ϕ\phi if

VρY(f)VρX(ϕ(f))V^{*}\rho_{Y}(f)V\sim\rho_{X}(\phi^{*}(f))

for every fC0(Y)f\in C_{0}(Y).

By Voiculescu’s theorem, every map ϕ\phi of the sort described in the definition is topologically covered by some isometry. If VV is a topological covering isometry for ϕ\phi then the *-homomorphism AdV:𝔹(HX)𝔹(HY)\text{Ad}_{V}\colon\mathbb{B}(H_{X})\to\mathbb{B}(H_{Y}) maps 𝔇(X)\mathfrak{D}^{*}(X) into 𝔇(Y)\mathfrak{D}^{*}(Y) and (X)\mathfrak{C}^{*}(X) into (Y)\mathfrak{C}^{*}(Y), and the induced maps on K-theory depend only on ϕ\phi and not on the chosen covering isometry ([3], chapter 5). Thus every continuous proper map ϕ:𝒰Y\phi\colon\mathcal{U}\to Y on an open subset of XX induces a map ϕ=(AdV):Kp(X)Kp(Y)\phi_{*}=(\text{Ad}_{V})_{*}\colon K_{p}(X)\to K_{p}(Y) on K-homology defined using a topological covering isometry.

2.2. Coarse K-theory

The target of the assembly map is the K-theory of the coarse algebra of the space XX, so we now review the definition and basic properties of this C*-algebra. Along the way we will also define the structure algebra.

For much of what follows we will be interested in proper metric spaces, i.e. a metric spaces whose closed and bounded subsets are all compact. Any complete Riemannian manifold is a proper metric space by the Hopf-Rinow theorem; indeed, this is one our main sources of examples.

Definition 2.5.

Let XX and YY be proper metric space and let ρX:C0(X)𝔹(HX)\rho_{X}\colon C_{0}(X)\to\mathbb{B}(H_{X}) and ρY:C0(Y)𝔹(HY)\rho_{Y}\colon C_{0}(Y)\to\mathbb{B}(H_{Y}) be representations. The support of a bounded operator T:HXHYT\colon H_{X}\to H_{Y} is the set Supp(T)\text{Supp}(T) of all points (y,x)Y×X(y,x)\in Y\times X with the following property: for every open neighborhood 𝒱×𝒰Y×X\mathcal{V}\times\mathcal{U}\subseteq Y\times X of (y,x)(y,x) there exist functions f1C0(𝒰)f_{1}\in C_{0}(\mathcal{U}) and f2C0(𝒱)f_{2}\in C_{0}(\mathcal{V}) such that ρY(f2)TρX(f1)0\rho_{Y}(f_{2})T\rho_{X}(f_{1})\neq 0. An operator in 𝔹(HX)\mathbb{B}(H_{X}) is said to be controlled if its support lies within a uniformly bounded neighborhood of the diagonal in X×XX\times X.

The set of all controlled operators forms a *-subalgebra, and we will be interested in the intersection of this *-subalgebra with the dual algebra and the locally compact algebra. Most of the results of this paper generalize easily to account for free and proper group actions, so we will include them at the outset (though we will not need to consider equivariant dual algebras or equivariant K-homology). For simplicity, the reader is invited to ignore the group action during the first reading.

So let XX be a proper metric space equipped with a free and proper action of a countable discrete group GG of isometries of XX. We will need to consider representations of C0(X)C_{0}(X) which are compatible with the GG-action; the motivating example is the representation of C0(X)C_{0}(X) on L2(X,μ)L^{2}(X,\mu) by multiplication operators where μ\mu is a GG-invariant Borel measure. More abstractly:

Definition 2.6.

Let HH be a Hilbert space equipped with a representation

ρ:C0(X)𝔹(H)\rho\colon C_{0}(X)\to\mathbb{B}(H)

and a unitary representation

U:G𝔹(H)U\colon G\to\mathbb{B}(H)

We say that the the triple (H,U,ρ)(H,U,\rho) is a GG-equivariant XX-module or simply a (X,G)(X,G)-module if Uγρ(f)=ρ(γf)UγU_{\gamma}\rho(f)=\rho(\gamma^{*}f)U_{\gamma} for every γG\gamma\in G, fC0(X)f\in C_{0}(X).

As usual if HH carries a unitary representation UU of GG then we say that an operator T𝔹(H)T\in\mathbb{B}(H) is GG-equivariant if UγTUγ=TU_{\gamma}TU_{\gamma}^{*}=T for every γG\gamma\in G.

Definition 2.7.

Let XX be a proper metric space equipped with a free and proper action of a countable discrete group of isometries and let (HX,U,ρX)(H_{X},U,\rho_{X}) be a GG-equivariant XX-module.

  • The structure algebra of the pair (X,G)(X,G) is the C*-algebra DG(X)D_{G}^{*}(X) obtained by taking the norm closure of the set of all GG-equivariant pseudolocal controlled operators.

  • The coarse algebra of the pair (X,G)(X,G) is the C*-ideal CG(X)C_{G}^{*}(X) in DG(X)D_{G}^{*}(X) obtained by taking the norm closure of the set of all GG-equivariant locally compact controlled operators.

  • Let QG(X)Q_{G}^{*}(X) denote the quotient C*-algebra DG(X)/CG(X)D_{G}^{*}(X)/C_{G}^{*}(X).

Remark 2.8.

It is important that we restrict to GG-invariant operators before passing to the closure rather than taking the GG-invariant part of the closure. Recent examples show that these two procedures do not in general yield the same C*-algebras.

As before, both DG(X)D_{G}^{*}(X) and CG(X)C_{G}^{*}(X) depend on the representation used to define them, but the ambiguity disappears at the level of K-theory under mild additional hypotheses. For the coarse algebra it is enough to assume that the Hilbert space is separable and the representation is ample, but for the structure algebra we must assume that the Hilbert space is separable and the representation is the countable direct sum of a fixed ample representation. Following [3], we refer to such representations as very ample.

There is a considerable amount of literature on CG(X)C_{G}^{*}(X) due in part to its relationship with index theory. Its K-theory groups are functorial for a class of maps between metric spaces called coarse maps which are compatible with large scale geometry (in the same way that continuous maps are compatible with small scale geometry). There is a general theory of abstract coarse structures, but for our present purposes it will suffice to review the basic features of the coarse structure of a metric space.

If SS is any set and α1,α2:SX\alpha_{1},\alpha_{2}\colon S\to X are maps, we say that α1\alpha_{1} and α2\alpha_{2} are close if there is a constant CC such that d(α1(s),α2(s))Cd(\alpha_{1}(s),\alpha_{2}(s))\leq C for every sSs\in S. A map ϕ:XY\phi\colon X\to Y between metric spaces is said to be coarse if ϕ1(B)\phi^{-1}(B) is a bounded subset of XX whenever BB is a bounded subset of YY and ϕα1\phi\circ\alpha_{1} and ϕα2\phi\circ\alpha_{2} are close whenever α1\alpha_{1} and α2\alpha_{2} are close. ϕ\phi is said to be a coarse equivalence if there is another coarse map ψ:YX\psi\colon Y\to X such that ϕψ\phi\circ\psi and ψϕ\psi\circ\phi are close to the identity maps on YY and XX, respectively. For instance, the standard embedding of \mathbb{Z} into \mathbb{R} is a coarse equivalence.

As with K-homology, the functoriality of coarse K-theory can be implemented at the level of the coarse algebra using another notion of covering isometry.

Definition 2.9.

Let XX and YY be proper metric spaces equipped with representations ρX:C0(X)𝔹(HX)\rho_{X}\colon C_{0}(X)\to\mathbb{B}(H_{X}) and ρY:C0(Y)𝔹(HY)\rho_{Y}\colon C_{0}(Y)\to\mathbb{B}(H_{Y}). An isometry V:HXHYV\colon H_{X}\to H_{Y} coarsely covers a coarse map

ϕ:XY\phi\colon X\to Y

if π1\pi_{1} and ϕπ2\phi\circ\pi_{2} are close as maps Supp(V)Y×XY\text{Supp}(V)\subseteq Y\times X\to Y. If HXH_{X} and HYH_{Y} carry unitary representations of a group GG then we say that a coarse covering isometry VV is GG-equivariant if in addition VUgX=UgYVVU_{g}^{X}=U_{g}^{Y}V.

According to chapter 6 of [3], every coarse map ϕ\phi admits a coarse covering isometry VV so long as ρX\rho_{X} is nondegenerate and ρY\rho_{Y} is ample; a straightforward variation on that argument shows that if a discrete group GG acts freely and properly on XX and YY by isometries and ϕ\phi is GG-invariant then VV can be taken to be a GG-equivariant coarse covering isometry. Moreover AdV\text{Ad}_{V} maps CG(X)C_{G}^{*}(X) into CG(Y)C_{G}^{*}(Y) and the induced map on K-theory is independent of the GG-equivariant coarse covering isometry chosen. Also note that if VV coarsely covers ϕ\phi then it also coarsely covers any map which is close to ϕ\phi, so ϕ\phi induces a map ϕ=(AdV):Kp(CG(X))Kp(CG(Y))\phi_{*}=(\text{Ad}_{V})_{*}\colon K_{p}(C_{G}^{*}(X))\to K_{p}(C_{G}^{*}(Y)) which depends only on the closeness equivalence class of ϕ\phi.

2.3. The Analytic Structure Group

Recall that we defined the structure algebra of a proper GG-space XX to be the C*-algebra DG(X)D_{G}^{*}(X) obtained by taking the norm closure of the space of GG-invariant pseudolocal controlled operators associated to a representation of C0(X)C_{0}(X).

Definition 2.10.

Let XX be a proper metric space equipped with a free and proper action of a discrete group GG by isometries and an ample (X,G)(X,G)-module. The analytic structure group of the pair (X,G)(X,G) is defined to be Sp(X,G)=K1p(DG(X))S_{p}(X,G)=K_{1-p}(D_{G}^{*}(X)).

While Kp(X)K_{p}(X) is an invariant of the small scale geometry of XX and Kp(CG(X))K_{p}(C_{G}^{*}(X)) is an invariant of the large scale geometry of XX, the analytic structure group depends simultaneously on both large and small scale behavior. As usual this will be clarified by understanding its functorial properties: we will prove that Sp(X,G)S_{p}(X,G) is functorial for uniform maps, i.e. maps which are both continuous and coarse. This functoriality is once again implemented using covering isometries.

Definition 2.11.

Let HXH_{X} and HYH_{Y} be Hilbert spaces which carry representations of C0(X)C_{0}(X) and C0(Y)C_{0}(Y), respectively. An isometry V:HXHYV\colon H_{X}\to H_{Y} uniformly covers a uniform map ϕ:XY\phi\colon X\to Y if it both topologically and coarsely covers ϕ\phi.

If XX and YY carry GG-actions and HXH_{X} and HYH_{Y} carry GG-equivariant representations then one can define GG-equivariant uniform covering isometries in the same way as in our discussion of GG-equivariant coarse covering isometries. GG-equivariant uniform covering isometries always exist, but the proof of this fact does not appear in the literature and thus we provide it here. Our strategy is to build ”local” covering isometries and patch them together using a partition of unity.

Lemma 2.12.

Let X=𝒰×G𝒰X=\mathcal{U}\times G\to\mathcal{U} and Y=𝒱×G𝒱Y=\mathcal{V}\times G\to\mathcal{V} be trivial GG-covers of proper metric spaces 𝒰\mathcal{U} and 𝒱\mathcal{V} and let HXH_{X} and HYH_{Y} be ample (X,G)(X,G)- and (Y,G)(Y,G)-modules, respectively. Then any continuous GG-equivariant map ϕ:XY\phi\colon X\to Y is topologically covered by a GG-equivariant isometry HXHYH_{X}\to H_{Y}.

Proof.

Let H𝒰H_{\mathcal{U}} denote the closure of C0(𝒰×{e})HXC_{0}(\mathcal{U}\times\{e\})H_{X} where eGe\in G is the identity and define H𝒱H_{\mathcal{V}} similarly. H𝒰H_{\mathcal{U}} and H𝒱H_{\mathcal{V}} carry ample representations of C0(𝒰)C_{0}(\mathcal{U}) and C0(𝒱)C_{0}(\mathcal{V}), respectively, and hence there is an isometry V:H𝒰H𝒱V\colon H_{\mathcal{U}}\to H_{\mathcal{V}} which topologically covers the restriction of ϕ\phi to U×{e}U\times\{e\}. We have that HX2(G)H𝒰H_{X}\cong\ell^{2}(G)\otimes H_{\mathcal{U}} and HY2(G)H𝒱H_{Y}\cong\ell^{2}(G)\otimes H_{\mathcal{V}}, so 1V:HXHY1\otimes V\colon H_{X}\to H_{Y} is an equivariant topological covering isometry for ϕ\phi. ∎

To construct GG-equivariant uniform covering isometries, we break XX and YY up into pieces for which the lemma applies. To make the argument work it is important that the (Y,G)(Y,G)-module be very ample, meaning it is the countable direct sum of a fixed ample representation.

Proposition 2.13.

Let HXH_{X} be an ample (X,G)(X,G)-module and let HYH_{Y} be a very ample (Y,G)(Y,G)-module. Then every equivariant uniform map ϕ:XY\phi\colon X\to Y is uniformly covered by an equivariant isometry V:HXHYV\colon H_{X}\to H_{Y}.

Proof.

Assume HY=HH_{Y}=\bigoplus_{\mathbb{N}}H where HH is a Hilbert space carrying a fixed ample representation ρY:C0(Y)𝔹(H)\rho_{Y}\colon C_{0}(Y)\to\mathbb{B}(H). Let πX:XXG\pi_{X}\colon X\to X_{G} and πY:YYG\pi_{Y}\colon Y\to Y_{G} be the natural quotient maps for the GG action, and choose countable locally finite open covers {𝒰m}\{\mathcal{U}_{m}\} of XX and {𝒱n}\{\mathcal{V}_{n}\} of YY with the following properties:

  • Each 𝒰m\mathcal{U}_{m} is GG-invariant and evenly covers πX(𝒰m)\pi_{X}(\mathcal{U}_{m}), and similarly for 𝒱n\mathcal{V}_{n}.

  • {πX(𝒰m)}\{\pi_{X}(\mathcal{U}_{m})\} and {πY(𝒱n)}\{\pi_{Y}(\mathcal{V}_{n})\} have uniformly bounded diameters

  • For every mm there exists n(m)n(m) such that ϕ(𝒰m)𝒱n(m)\phi(\mathcal{U}_{m})\subseteq\mathcal{V}_{n(m)}

Let HmXH_{m}^{X} denote the closure of ρX(C0(𝒰m))HX\rho_{X}(C_{0}(\mathcal{U}_{m}))H_{X} and let HnYH_{n}^{Y} denote the closure of ρY(C0(𝒱n))H\rho_{Y}(C_{0}(\mathcal{V}_{n}))H. Thus HmXH_{m}^{X} carries an ample GG-equivariant representation of C0(𝒰m)C_{0}(\mathcal{U}_{m}) and HnYH_{n}^{Y} carries an ample GG-equivariant representation of C0(𝒱n)C_{0}(\mathcal{V}_{n}). Since ϕ\phi restricts to a continuous map 𝒰m𝒱n(m)\mathcal{U}_{m}\to\mathcal{V}_{n(m)}, there is a GG-equivariant isometry Vm:HmXHn(m)YV_{m}\colon H_{m}^{X}\to H_{n(m)}^{Y} which topologically covers ϕ|𝒰m\phi|_{\mathcal{U}_{m}} by Lemma 2.12.

Let {hm}\{h_{m}\} be a GG-invariant partition of unity subordinate to the open cover {𝒰m}\{\mathcal{U}_{m}\} and define V:HXHY=HV\colon H_{X}\to H_{Y}=\bigoplus_{\mathbb{N}}H to be the strong limit

V=mVmρX(hm1/2)V=\bigoplus_{m}V_{m}\rho_{X}(h_{m}^{1/2})

It is clear that VV is a GG-equivariant isometry. To show that VV topologically covers ϕ\phi, we must show that ρX(gϕ)VρY(g)V\rho_{X}(g\circ\phi)\sim V^{*}\rho_{Y}(g)V for every gC0(Y)g\in C_{0}(Y). We have

VρY(g)V\displaystyle V^{*}\rho_{Y}(g)V =mρX(hm1/2)VmρY(g)VmρX(hm1/2)\displaystyle=\bigoplus_{m}\rho_{X}(h_{m}^{1/2})V_{m}^{*}\rho_{Y}(g)V_{m}\rho_{X}(h_{m}^{1/2})
mρX(hm1/2)ρX(gϕ|𝒰m)ρX(hm1/2)\displaystyle\sim\sum_{m}\rho_{X}(h_{m}^{1/2})\rho_{X}(g\circ\phi|_{\mathcal{U}_{m}})\rho_{X}(h_{m}^{1/2})
=gϕ\displaystyle=g\circ\phi

since VmV_{m} topologically covers ϕ|𝒰m\phi|_{\mathcal{U}_{m}}.

Finally, to show that VV coarsely covers ϕ\phi note that Supp(V)=mSupp(Vm)\text{Supp}(V)=\bigcup_{m}\text{Supp}(V_{m}) and Supp(Vm)𝒱n(m)×𝒰m\text{Supp}(V_{m})\subseteq\mathcal{V}_{n(m)}\times\mathcal{U}_{m}. Since the diameters of the sets 𝒱n\mathcal{V}_{n} and 𝒰m\mathcal{U}_{m} are uniformly bounded and ϕ\phi maps 𝒰m\mathcal{U}_{m} into 𝒱n(m)\mathcal{V}_{n(m)}, the restrictions of π1\pi_{1} and ϕπ2\phi\circ\pi_{2} to Supp(V)\text{Supp}(V) are close. This completes the proof. ∎

If VV is a GG-equivariant covering isometry for a GG-equivariant uniform map ϕ\phi then AdV\text{Ad}_{V} maps DG(X)D_{G}^{*}(X) into DG(Y)D_{G}^{*}(Y) and the induced map on K-theory is independent of the GG-equivariant covering isometry chosen. These observations follow immediately by combining the corresponding arguments for K-homology and coarse K-theory. The conclusion is that every GG-equivariant uniform map ϕ:XY\phi\colon X\to Y induces a map ϕ=(AdV):Sp(X,G)Sp(Y,G)\phi_{*}=(\text{Ad}_{V})_{*}\colon S_{p}(X,G)\to S_{p}(Y,G).

2.4. The Assembly Map

We are now ready to construct the (equivariant) coarse assembly map described in the introduction. As usual let XX be a proper metric space on which GG acts freely and properly by isometries. Let XGX_{G} denote the quotient space of XX by GG. There is an isomorphism

(2.1) QG(X)𝔔(XG)Q_{G}^{*}(X)\cong\mathfrak{Q}^{*}(X_{G})

and hence the long exact sequence in K-theory associated to the short exact sequence

0CG(X)DG(X)QG(X)00\to C_{G}^{*}(X)\to D_{G}^{*}(X)\to Q_{G}^{*}(X)\to 0

takes the form:

(2.2) Kp+1(CG(X))Sp(X,G)Kp(XG)Kp(CG(X))\to K_{p+1}(C_{G}^{*}(X))\to S_{p}(X,G)\to K_{p}(X_{G})\to K_{p}(C_{G}^{*}(X))\to

In the case where XGX_{G} is a compact manifold and XX is its universal cover, this is the analytic surgery exact sequence investigated by Higson and Roe in [4], [5], and [6]. In any event, the boundary map Kp(XG)Kp(CG(X))K_{p}(X_{G})\to K_{p}(C_{G}^{*}(X)) is the desired coarse assembly map.

The key is the isomorphism (2.1). It is proved in [3] and in [9] (using sheaf-theoretic techniques). We will give a brief outline of the argument in [3] so that we can adapt it to the relative setting in the next section. The main tool is a certain truncation operator defined as follows:

Definition 2.14.

Let XX be a proper metric space and let ρ:C0(X)𝔹(HX)\rho\colon C_{0}(X)\to\mathbb{B}(H_{X}) be a representation. Let {Un}\{U_{n}\} be a countable locally finite collection of open subsets of XX and let {hn}\{h_{n}\} be a subordinate partition of unity. Given any T𝔹(HX)T\in\mathbb{B}(H_{X}), define 𝔗(T)\mathfrak{T}(T) to be the strong limit of the series nρ(hn1/2)Tρ(hn1/2)\sum_{n}\rho(h_{n}^{1/2})T\rho(h_{n}^{1/2}). 𝔗\mathfrak{T} is called the truncation of TT relative to the open cover {𝒰n}\{\mathcal{U}_{n}\}.

Here are the properties of the truncation construction which we will need:

Proposition 2.15.

In the setting of Definition 2.14, the following hold:

  • 𝔗:𝔹(HX)𝔹(HX)\mathfrak{T}\colon\mathbb{B}(H_{X})\to\mathbb{B}(H_{X}) is a continuous linear map.

  • Supp(𝔗(T))n𝒰n×𝒰n¯\text{Supp}(\mathfrak{T}(T))\subseteq\bigcup_{n}\overline{\mathcal{U}_{n}\times\mathcal{U}_{n}}.

  • If TT is pseudolocal then 𝔗(T)\mathfrak{T}(T) is pseudolocal and T𝔗(T)T-\mathfrak{T}(T) is locally compact.

Proof.

See Chapter 12 of [3]. ∎

The isomorphism 2.1 can be constructed using truncation operators in two steps. The first step is to pass from equivariant operators associated to XX to ordinary operators on XGX_{G}. This is more or less straightforward for operators supported in an open subset of XX of the form 𝒰=𝒰G×G\mathcal{U}=\mathcal{U}_{G}\times G where 𝒰G\mathcal{U}_{G} is an open subset of XGX_{G}, so we cover XX by countably many evenly covered open sets of uniformly bounded diameter, lift to an open cover {𝒰n}\{\mathcal{U}_{n}\} of XX, and truncate a general GG-equivariant operator TT along this open cover using a GG-invariant partition of unity. 𝔗(T)\mathfrak{T}(T) descends to a pseudolocal controlled operator associated to XGX_{G} and it differs from TT by a locally compact GG-equivariant controlled operator by Proposition 2.15. This yields an isomorphism QG(X)Q(XG)Q_{G}^{*}(X)\cong Q^{*}(X_{G}).

The second step is to pass from a general pseudolocal operator TT associated to XGX_{G} to a pseudolocal controlled operator. Once again, the strategy is to truncate TT along a uniformly bounded open cover of XGX_{G}; Proposition 2.15 guarantees that 𝔗(T)\mathfrak{T}(T) is pseudolocal and controlled and that it differs from TT by a locally compact operator. This yields an isomorphism Q(XG)𝔔(XG)Q^{*}(X_{G})\cong\mathfrak{Q}^{*}(X_{G}). The fact that 𝔗\mathfrak{T} is continuous is used in both steps to handle the fact that a general operator in the coarse algebra or the structure algebra is the norm limit of controlled operators but may not itself be controlled.

3. Mayer-Vietoris Sequences

In this section we construct Mayer-Vietoris sequences for K-homology, coarse K-theory, and the analytic structure group using an abstract Mayer-Vietoris sequence for C*-algebra K-theory groups. We also provide formulas for the Mayer-Vietoris boundary maps; these will be used in the next section to prove the partitioned manifold index theorem. The Mayer-Vietoris sequence for K-homology follows from standard results in classical algebraic topology and the Mayer-Vietoris sequence for coarse K-theory was described in [7], but the author is unaware of any previous discussion of the Mayer-Vietoris sequence for the analytic structure group.

3.1. The Abstract Mayer-Vietoris Sequence

Let AA be a C*-algebra and let I1I_{1} and I2I_{2} be C*-ideals in AA such that I1+I2=AI_{1}+I_{2}=A. There is a Mayer-Vietoris sequence in K-theory associated to these data which takes the form:

(3.1) Kp+1(A)Kp(I1I2)Kp(I1)Kp(I2)Kp(A)\to K_{p+1}(A)\to K_{p}(I_{1}\cap I_{2})\to K_{p}(I_{1})\oplus K_{p}(I_{2})\to K_{p}(A)\to

This is apparently a folklore result among operator algebraists; it appears in [7] but it is probably older. To construct it, consider the C*-algebra Ω(A,I1,I2)\Omega(A,I_{1},I_{2}) defined to be the set of all continuous paths f:[0,1]Af\colon[0,1]\to A such that f(0)I1f(0)\in I_{1} and f(1)I2f(1)\in I_{2}. There is a short exact sequence

(3.2) 0SAΩ(A,I2,I2)I1I200\to SA\to\Omega(A,I_{2},I_{2})\to I_{1}\oplus I_{2}\to 0

where SASA is the suspension of AA, the first map is inclusion, and the second map is evaluation at 0 and 11.

Lemma 3.1.

The *-homomorphism I1I2Ω(A,I2,I2)I_{1}\cap I_{2}\to\Omega(A,I_{2},I_{2}) which sends aI1I2a\in I_{1}\cap I_{2} to the constant path based at aa induces an isomorphism in K-theory.

Proof.

The space C[0,1]I1I2C[0,1]\otimes I_{1}\cap I_{2} of continuous paths in I1I2I_{1}\cap I_{2} is an ideal in Ω(A,I1,I2)\Omega(A,I_{1},I_{2}) which is homotopy equivalent to I1I2I_{1}\cap I_{2}. Thus it suffices to show that the inclusion C[0,1]I1I2C[0,1]\otimes I_{1}\cap I_{2} in Ω(A,I1,I2)\Omega(A,I_{1},I_{2}) induces an isomorphism on K-theory. The quotient QQ of Ω(A,I1,I2)\Omega(A,I_{1},I_{2}) by this ideal is the C*-algebra of paths g:[0,1]A/(I1I2)g\colon[0,1]\to A/(I_{1}\cap I_{2}) such that g(0)I1/(I1I2)g(0)\in I_{1}/(I_{1}\cap I_{2}) and g(1)I2/(I1I2)g(1)\in I_{2}/(I_{1}\cap I_{2}). Since

A/(I1I2)I1/(I1I2)I2/(I1I2)A/(I_{1}\cap I_{2})\cong I_{1}/(I_{1}\cap I_{2})\oplus I_{2}/(I_{1}\cap I_{2})

(this is where A=I1+I2A=I_{1}+I_{2} is used), it follows that

Q(C0[0,1)I1/(I1I2))(C0(0,1]I2/(I1I2))Q\cong\left(C_{0}[0,1)\otimes I_{1}/(I_{1}\cap I_{2})\right)\oplus\left(C_{0}(0,1]\otimes I_{2}/(I_{1}\cap I_{2})\right)

Thus QQ has trivial K-theory and the proof is complete. ∎

The Mayer Vietoris sequence (3.1) can thus be obtained as the long exact sequence in K-theory associated to the short exact sequence (3.2)

We conclude by providing formulas for the boundary maps in the Mayer-Vietoris sequence. These formulas use the following device:

Definition 3.2.

Let AA be a unital C*-algebra and assume A=I1+I2A=I_{1}+I_{2} where I1I_{1} and I2I_{2} are closed C*-ideals in AA. A partition of unity for this decomposition is a pair (a1,a2)(a_{1},a_{2}) where ajIja_{j}\in I_{j} are positive elements of AA satisfying a12+a22=1a_{1}^{2}+a_{2}^{2}=1.

Partitions of unity exist and are unique up to homotopy by a straightforward functional calculus argument. Let us use a partition of unity to calculate the Mayer-Vietoris boundary map MV:K1(A)K0(I1I2)\partial_{MV}\colon K_{1}(A)\to K_{0}(I_{1}\cap I_{2}). This begins with the following lemma.

Lemma 3.3.

Let uMn(A)\textbf{u}\in M_{n}(A) be a unitary representing a class in K1(A)K_{1}(A) and define v=a1u+a21\textbf{v}=a_{1}\textbf{u}+a_{2}\textbf{1}. We have:

  • v1\textbf{v}\sim\textbf{1} modulo Mn(I1)M_{n}(I_{1})

  • vu\textbf{v}\sim\textbf{u} modulo Mn(I2)M_{n}(I_{2})

  • v is a unitary modulo Mn(I1I2)M_{n}(I_{1}\cap I_{2})

Proof.

Since a1I1a_{1}\in I_{1} and a2I2a_{2}\in I_{2} are positive elements satisfying a12+a22=1{a_{1}}^{2}+{a_{2}}^{2}=1, it follows that a2=1a12a_{2}=\sqrt{1-{a_{1}}^{2}} and hence a21a_{2}\sim 1 modulo I1I_{1}. Thus a1u+a211a_{1}\textbf{u}+a_{2}\textbf{1}\sim\textbf{1} modulo Mn(I1)M_{n}(I_{1}), and vu\textbf{v}\sim\textbf{u} modulo Mn(I2)M_{n}(I_{2}) follows similarly. To see that v is a unitary modulo Mn(I1I2)M_{n}(I_{1}\cap I_{2}), we use the fact that a1a_{1} and a2a_{2} commute with any element of AA modulo I1I2I_{1}\cap I_{2} since a1a21a_{1}\sim a_{2}\sim 1. Thus:

vv\displaystyle\textbf{v}\textbf{v}^{*} =(a1u+a21)(ua1+a21)\displaystyle=(a_{1}\textbf{u}+a_{2}\textbf{1})(\textbf{u}^{*}a_{1}+a_{2}\textbf{1})
=(a12+a22)1+a1ua2+a2ua1\displaystyle=({a_{1}}^{2}+{a_{2}}^{2})\textbf{1}+a_{1}\textbf{u}a_{2}+a_{2}\textbf{u}^{*}a_{1}
1+a1a2u+a1a2u\displaystyle\sim\textbf{1}+a_{1}a_{2}\textbf{u}+a_{1}a_{2}\textbf{u}^{*}
1 modulo Mn(I1I2)\displaystyle\sim\textbf{1}\text{ modulo $M_{n}(I_{1}\cap I_{2})$}

Similarly:

vv\displaystyle\textbf{v}^{*}\textbf{v} =(ua1+a21)(a1u+a21)\displaystyle=(\textbf{u}^{*}a_{1}+a_{2}\textbf{1})(a_{1}\textbf{u}+a_{2}\textbf{1})
(a12+a22)1+ua1a2+a2a1u\displaystyle\sim({a_{1}}^{2}+{a_{2}}^{2})\textbf{1}+\textbf{u}^{*}a_{1}a_{2}+a_{2}a_{1}\textbf{u}
1+a1a2u+a1a2u\displaystyle\sim\textbf{1}+a_{1}a_{2}\textbf{u}^{*}+a_{1}a_{2}\textbf{u}
1 modulo Mn(I1I2)\displaystyle\sim\textbf{1}\text{ modulo $M_{n}(I_{1}\cap I_{2})$}

Our formula for MV\partial_{MV} is as follows:

Proposition 3.4.

Let u and v be as in the previous lemma. Then MV[u]=[v]\partial_{MV}[\textbf{u}]=\partial[\textbf{v}] where [v][\textbf{v}] is the class of v in K1(A/(I1I2))K_{1}(A/(I_{1}\cap I_{2})) and \partial is the boundary map for the short exact sequence

0I1I2AA/(I1I2)00\to I_{1}\cap I_{2}\to A\to A/(I_{1}\cap I_{2})\to 0

The proof uses an explicit formula for the K-theory boundary map K1(A/(I1I2))K0(I1I2)K_{1}(A/(I_{1}\cap I_{2}))\to K_{0}(I_{1}\cap I_{2}) and some elementary homotopies. Further detail is available in [10].

There is a similar formula for the Mayer-Vietoris boundary map in the other degree, but in this case we must assume that A=I1+I2A=I_{1}+I_{2} admits a partition of unity (a1,a2)(a_{1},a_{2}) such that a1a_{1} and a2a_{2} are projections; this is required to ensure that a1p+a2a_{1}\textbf{p}+a_{2} is a projection if p is a projection. Partitions of unity of this sort are available in the examples relevant to this paper.

Proposition 3.5.

Assume A=I1+I2A=I_{1}+I_{2} admits a partition of unity (a1,a2)(a_{1},a_{2}) consisting of projections. Then the Mayer-Vietoris boundary map MV:K0(A)K1(I1I2)\partial_{MV}:K_{0}(A)\to K_{1}(I_{1}\cap I_{2}) satisfies MV[p]=[a1p+a2]\partial_{MV}[\textbf{p}]=\partial[a_{1}\textbf{p}+a_{2}] where \partial is the boundary map associated to the short exact sequence

0I1I2AA/I1I200\to I_{1}\cap I_{2}\to A\to A/I_{1}\cap I_{2}\to 0

This formula can be reduced to the formula in the other degree using suspensions and Bott periodicity. Again, the details are provided in [10].

3.2. The Mayer-Vietoris and Analytic Surgery Diagram

Let XX be a proper metric space and let Y1Y_{1} and Y2Y_{2} be subspaces such that X=Y1Y2X=Y_{1}\cup Y_{2}. In this section we will associate to Y1Y_{1} and Y2Y_{2} ideals 𝔔(Y1X)\mathfrak{Q}^{*}(Y_{1}\subseteq X) and 𝔔(Y2X)\mathfrak{Q}^{*}(Y_{2}\subseteq X) in 𝔇(X)\mathfrak{D}^{*}(X) and show that 𝔇(X)=𝔇(Y1X)+𝔇(Y2X)\mathfrak{D}^{*}(X)=\mathfrak{D}^{*}(Y_{1}\subseteq X)+\mathfrak{D}^{*}(Y_{2}\subseteq X). Combined with some excision results, the abstract Mayer-Vietoris sequence of the previous section will yield a Mayer-Vietoris sequence in K-homology:

Kp+1(X)Kp(Y1Y2)Kp(Y1)Kp(Y2)Kp(X)\to K_{p+1}(X)\to K_{p}(Y_{1}\cap Y_{2})\to K_{p}(Y_{1})\oplus K_{p}(Y_{2})\to K_{p}(X)\to

We will also carry out similar constructions for the coarse K-theory group and the analytic structure group of XX and conclude by showing that the three Mayer-Vietoris sequences can be interwoven with the analytic surgery exact sequence to form the braid diagram which appears in the introduction.

To begin, fix a representation ρ\rho of C0(X)C_{0}(X) on a separable Hilbert space HH and let YXY\subseteq X be a subspace. Say that an operator T𝔹(H)T\in\mathbb{B}(H) is locally compact for XYX-Y if ρ(f)TTρ(f)0\rho(f)T\sim T\rho(f)\sim 0 for every fC0(XY)f\in C_{0}(X-Y), and say that TT is supported near YY if Supp(T)BR(Y)×BR(Y)\text{Supp}(T)\subseteq B_{R}(Y)\times B_{R}(Y) for some RR-neighborhood BR(Y)B_{R}(Y).

Definition 3.6.

Let XX be a proper metric space equipped with a free, proper, and isometric GG-action and let YY be a GG-invariant subspace.

  • Let 𝔇(YX)\mathfrak{D}^{*}(Y\subseteq X) denote the C*-ideal in 𝔇(X)\mathfrak{D}^{*}(X) consisting of pseudolocal operators which are locally compact for XYX-Y. Let 𝔔(YX)\mathfrak{Q}^{*}(Y\subseteq X) denote the ideal 𝔇(YX)/(X)\mathfrak{D}^{*}(Y\subseteq X)/\mathfrak{C}^{*}(X) in 𝔔(X)\mathfrak{Q}^{*}(X).

  • Let CG(YX)C_{G}^{*}(Y\subseteq X) denote the C*-ideal in CG(X)C_{G}^{*}(X) obtained by taking the norm closure of the set of all GG-invariant locally compact controlled operators which are supported near YY.

  • Let DG(YX)D_{G}^{*}(Y\subseteq X) denote the C*-ideal in DG(X)D_{G}^{*}(X) obtained by taking the norm closure of the set of all GG-invariant pseudolocal controlled operators which are locally compact for XYX-Y and supported near YY.

Remark 3.7.

As before, we do not need to work equivariantly with the dual algebra.

Defining QG(YX)=DG(YX)/CG(YX)Q_{G}^{*}(Y\subseteq X)=D_{G}^{*}(Y\subseteq X)/C_{G}^{*}(Y\subseteq X), there is still an isomorphism

QG(YX)𝔔(YGXG)Q_{G}^{*}(Y\subseteq X)\cong\mathfrak{Q}^{*}(Y_{G}\subseteq X_{G})

The proof proceeds exactly as before, only now we build the required truncation operator using a collection of open sets {𝒰n}\{\mathcal{U}_{n}\} in XX such that Yn𝒰nBR(Y)Y\subseteq\bigcup_{n}\mathcal{U}_{n}\subseteq B_{R}(Y) for some R>0R>0. This ensures that 𝔗(T)\mathfrak{T}(T) is supported near YY for any operator TT.

An crucial fact about these ideals is that their K-theory depends only on YY.

Proposition 3.8.

Let XX and YY be as in the previous definition and fix an ample representation of C0(X)C_{0}(X) on a separable Hilbert space. Then the inclusion i:YXi\colon Y\hookrightarrow X is continuous, coarse, and GG-invariant, and we have:

  • If YY is closed then ii induces an isomorphism

    Kp(𝔔(Y))Kp(𝔔(YX))K_{p}(\mathfrak{Q}^{*}(Y))\cong K_{p}(\mathfrak{Q}^{*}(Y\subseteq X))
  • ii induces an isomorphism

    Kp(CG(Y))Kp(CG(YX))K_{p}(C_{G}*(Y))\cong K_{p}(C_{G}^{*}(Y\subseteq X))
  • If YY is closed and the representation is very ample then ii induces an isomorphism

    Kp(DG(Y))Kp(DG(YX))K_{p}(D_{G}^{*}(Y))\cong K_{p}(D_{G}^{*}(Y\subseteq X))
Proof.

For K-homology this is proved in Chapter 5 of [3] and for coarse K-theory this is proved in [7]. For the structure group we proceed as follows. ii is an equivariant uniform map, so it is uniformly covered by an equivariant isometry V:HYHXV\colon H_{Y}\to H_{X}. VV is in particular an equivariant coarse covering isometry, and following the proof of Proposition 2.13 it is the lift of a topological covering isometry VGV_{G} for the inclusion YGXGY_{G}\hookrightarrow X_{G}. Thus there is a commutative diagram:

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}CG(Y)\textstyle{C_{G}^{*}(Y)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ad(V)\scriptstyle{\text{Ad}(V)}DG(Y)\textstyle{D_{G}^{*}(Y)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ad(V)\scriptstyle{\text{Ad}(V)}DG(Y)/CG(Y)\textstyle{D_{G}^{*}(Y)/C_{G}^{*}(Y)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}0\textstyle{0}𝔇(YG)/(YG)\textstyle{\mathfrak{D}^{*}(Y_{G})/\mathfrak{C}^{*}(Y_{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ad(VG)\scriptstyle{\text{Ad}(V_{G})}𝔇(YGXG)/(YGXG)\textstyle{\mathfrak{D}^{*}(Y_{G}\subseteq X_{G})/\mathfrak{C}^{*}(Y_{G}\subseteq X_{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}CG(YX)\textstyle{C_{G}^{*}(Y\subseteq X)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}DG(YX)\textstyle{D_{G}^{*}(Y\subseteq X)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}DG(YX)/CG(Yx)\textstyle{D_{G}^{*}(Y\subseteq X)/C_{G}^{*}(Y\subseteq x)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

This gives rise to a commutative diagram in K-theory:

Kp+1(YG)\textstyle{K_{p+1}(Y_{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Kp+1(CG(Y))\textstyle{K_{p+1}(C_{G}^{*}(Y))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Kp+1(DG(Y))\textstyle{K_{p+1}(D_{G}^{*}(Y))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Kp(YG)\textstyle{K_{p}(Y_{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Kp(CG(Y))\textstyle{K_{p}(C_{G}^{*}(Y))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Kp+1(YGXG)\textstyle{K_{p+1}(Y_{G}\subseteq X_{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Kp+1(CG(YX))\textstyle{K_{p+1}(C_{G}^{*}(Y\subseteq X))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Kp+1(DG(YX))\textstyle{K_{p+1}(D_{G}^{*}(Y\subseteq X))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Kp(YGXG)\textstyle{K_{p}(Y_{G}\subseteq X_{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Kp(CG(YX))\textstyle{K_{p}(C_{G}^{*}(Y\subseteq X))}

We have proven that all of the vertical maps except the middle one are isomorphisms, so the middle map is an isomorphism by the five lemma. ∎

We are now ready to construct the desired Mayer-Vietoris sequences associated to a (GG-invariant) decomposition X=Y1Y2X=Y_{1}\cup Y_{2}. Let P1P_{1} and P2P_{2} denote the bounded Hilbert space operators corresponding to bounded Borel functions f1f_{1} and f2f_{2} which satisfy f1+f2=1f_{1}+f_{2}=1, f1|XY1=0f_{1}|_{X-Y_{1}}=0, and f2|XY2=0f_{2}|_{X-Y_{2}}=0 (such functions exist since XY1X-Y_{1} and XY2X-Y_{2} are disjoint open sets). P1P_{1} is a GG-invariant locally compact controlled operator which is supported near Y1Y_{1}, so in particular P1P_{1} is in 𝔇(Y1X)\mathfrak{D}^{*}(Y_{1}\subseteq X), CG(Y1X)C_{G}^{*}(Y_{1}\subseteq X), and DG(Y1X)D_{G}^{*}(Y_{1}\subseteq X) (and similarly for P2P_{2}). Moreover P1+P2=1P_{1}+P_{2}=1, so we have shown that:

  • 𝔔(X)=𝔔(Y1X)+𝔔(Y2X)\mathfrak{Q}^{*}(X)=\mathfrak{Q}^{*}(Y_{1}\subseteq X)+\mathfrak{Q}^{*}(Y_{2}\subseteq X)

  • CG(X)=CG(Y1X)+CG(Y2X)C_{G}^{*}(X)=C_{G}^{*}(Y_{1}\subseteq X)+C_{G}^{*}(Y_{2}\subseteq X)

  • DG(X)=DG(Y1X)+DG(Y2X)D_{G}^{*}(X)=D_{G}^{*}(Y_{1}\subseteq X)+D_{G}^{*}(Y_{2}\subseteq X)

This is the input required to form the abstract Mayer-Vietoris sequence (3.1), but to obtain a Mayer-Vietoris sequence which depends only on the geometry of the decomposition X=Y1Y2X=Y_{1}\cup Y_{2} we must check that the intersection of the ideals corresponding to Y1Y_{1} and Y2Y_{2} agrees with the ideal corresponding to Y1Y2Y_{1}\cap Y_{2}. This requires certain excision conditions which must be formulated seperately for the dual algebra, the coarse algebra, and the structure algebra.

For the coarse algebra and structure algebra the key condition is that Y1Y_{1} and Y2Y_{2} form an ω\omega-excisive pair, meaning for every R>0R>0 there exists S>0S>0 such that

BR(Y1)BR(Y2)BS(Y1Y2)B_{R}(Y_{1})\cap B_{R}(Y_{2})\subseteq B_{S}(Y_{1}\cap Y_{2})

For instance if X=X=\mathbb{R} with the standard metric, Y1=(,0]Y_{1}=(-\infty,0], and Y2=[0,)Y_{2}=[0,\infty) then Y1Y_{1} and Y2Y_{2} form a coarsely excisive pair. If on the other hand we equip \mathbb{R} with a metric obtained by embedding it into 2\mathbb{R}^{2} in such a way that the two rays remain within a bounded distance of each other then the decomposition is not ω\omega-excisive.

Proposition 3.9.

Let X=Y1Y2X=Y_{1}\cup Y_{2} be a GG-invariant decomposition.

  • If Y1Y_{1} and Y2Y_{2} are closed then

    𝔇(Y1X)𝔇(Y2X)=𝔇(Y1Y2X)\mathfrak{D}^{*}(Y_{1}\subseteq X)\cap\mathfrak{D}^{*}(Y_{2}\subseteq X)=\mathfrak{D}^{*}(Y_{1}\cap Y_{2}\subseteq X)
  • If Y1Y_{1} and Y2Y_{2} form an ω\omega-excisive pair then:

    CG(Y1X)CG(Y2X)=CG(Y1Y2X)C_{G}^{*}(Y_{1}\subseteq X)\cap C_{G}^{*}(Y_{2}\subseteq X)=C_{G}^{*}(Y_{1}\cap Y_{2}\subseteq X)
  • If Y1Y_{1} and Y2Y_{2} are closed and form an ω\omega-excisive pair then

    DG(Y1X)DG(Y2X)=DG(Y1Y2X)D_{G}^{*}(Y_{1}\subseteq X)\cap D_{G}^{*}(Y_{2}\subseteq X)=D_{G}^{*}(Y_{1}\cap Y_{2}\subseteq X)
Proof.

The proof of the identity for the coarse algebra appears in [7], and the identity for the structure algebra follows by combining the identity for the dual algebra with the identity for the coarse algebra. So we need only prove the identity for the dual algebra.

The containment “\subseteq” is trivial since C0(XYi)C0(XY1Y2)C_{0}(X-Y_{i})\subseteq C_{0}(X-Y_{1}\cap Y_{2}), so that an operator which is locally compact for XY1Y2X-Y_{1}\cap Y_{2} is automatically locally compact for XY1X-Y_{1} and XY2X-Y_{2}. To prove “\supseteq”, begin by observing that XY1Y2X-Y_{1}\cap Y_{2} is the disjoint union of XY1X-Y_{1} and XY2X-Y_{2}. So given any fC0(XY1Y2)f\in C_{0}(X-Y_{1}\cap Y_{2}) we have f=f1+f2f=f_{1}+f_{2} where fi=f|Yif_{i}=f|_{Y_{i}}. For any T𝔇(Y1X)𝔇(Y2X)T\in\mathfrak{D}^{*}(Y_{1}\subseteq X)\cap\mathfrak{D}^{*}(Y_{2}\subseteq X) we have that Tf1f1TTf2f2T0Tf_{1}\sim f_{1}T\sim Tf_{2}\sim f_{2}T\sim 0, so it follows that T(f1+f2)(f1+f2)T0T(f_{1}+f_{2})\sim(f_{1}+f_{2})T\sim 0. ∎

Now, let 𝛀(CG,X,Y1,Y2)\boldsymbol{\Omega}(C_{G}^{*},X,Y_{1},Y_{2}) denote the short exact sequence (3.2)\eqref{MVShortExact} whose long exact sequence in K-theory is the Mayer-Vietoris sequence for the decomposition CG(X)=CG(Y1X)+CG(Y2X)C_{G}^{*}(X)=C_{G}^{*}(Y_{1}\subseteq X)+C_{G}^{*}(Y_{2}\subseteq X), and define 𝛀(DG,X,Y1,Y2)\boldsymbol{\Omega}(D_{G}^{*},X,Y_{1},Y_{2}) and 𝛀(QG,X,Y1,Y2)\boldsymbol{\Omega}(Q_{G}^{*},X,Y_{1},Y_{2}) similarly. There is a complex

0𝛀(CG,X,Y1,Y2)𝛀(DG,X,Y1,Y2)𝛀(QG,X,Y1,Y2)00\to\boldsymbol{\Omega}(C_{G}^{*},X,Y_{1},Y_{2})\to\boldsymbol{\Omega}(D_{G}^{*},X,Y_{1},Y_{2})\to\boldsymbol{\Omega}(Q_{G}^{*},X,Y_{1},Y_{2})\to 0

Passing to K-theory the columns induce Mayer-Vietoris sequences and the rows induce analytic surgery exact sequences, so by the naturality properties of K-theory we obtain:

Kp((Y1)G)Kp((Y2)G)\textstyle{K_{p}((Y_{1})_{G})\oplus K_{p}((Y_{2})_{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Kp(CG(Y1))Kp(CG(Y2))\textstyle{K_{p}(C_{G}^{*}(Y_{1}))\oplus K_{p}(C_{G}^{*}(Y_{2}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Kp(CG(X))\textstyle{K_{p}(C_{G}^{*}(X))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Sp1(X,G)\textstyle{S_{p-1}(X,G)}Kp(CG(Y1Y2))\textstyle{K_{p}(C_{G}^{*}(Y_{1}\cap Y_{2}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Kp(XG)\textstyle{K_{p}(X_{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Sp1(Y1,G)Sp1(Y2,G)\textstyle{S_{p-1}(Y_{1},G)\oplus S_{p-1}(Y_{2},G)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Kp1(CG(Y1Y2))\textstyle{K_{p-1}(C_{G}^{*}(Y_{1}\cap Y_{2}))}Sp(X,G)\textstyle{S_{p}(X,G)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Sp1(Y1Y2,G)\textstyle{S_{p-1}(Y_{1}\cap Y_{2},G)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Kp1((Y1Y2)G)\textstyle{K_{p-1}((Y_{1}\cap Y_{2})_{G})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Kp1((Y1)G)Kp1((Y2)G)\textstyle{K_{p-1}((Y_{1})_{G})\oplus K_{p-1}((Y_{2})_{G})}

This is the Mayer-Vietoris and analytic surgery diagram.

4. Partitioned Manifolds

In this section we use the Mayer-Vietoris and analytic surgery diagram to reprove and generalize Roe’s index theorem for partitioned manifolds. Before formulating Roe’s theorem, let us take a moment to briefly review the relationship between index theory and the assembly map described in this paper.

4.1. Index Theory and Assembly

Let MM be a Riemannian manifold and let DD be a first order differential operator acting on smooth sections of a vector bundle SMS\to M. Recall that the symbol of DD is the bundle map σD:TMEnd(S)\sigma_{D}:T^{*}M\to\text{End}(S) obtained by freezing the coefficients of DD and passing to the Fourier transform of its top order part. DD is said to be elliptic if its symbol is invertible away from the zero section, and DD has finite propagation speed if the restriction of σD\sigma_{D} to the unit sphere bundle in TMT^{*}M is uniformly bounded in norm.

Definition 4.1.

A Dirac-type operator on a Riemannian manifold MM is a first order, symmetric, elliptic differential operator acting on smooth sections of a vector bundle SMS\to M which has finite propagation speed.

Other authors sometimes define Dirac-type operators more narrowly than we have here.

A bundle SMS\to M is said to be graded if it comes equipped with a decomposition S=S+SS=S^{+}\oplus S^{-} into subbundles. SS is pp-multigraded, p1p\geq 1, if it is graded and it comes equipped with pp odd-graded anti-commuting unitary operators ε1,,εp\varepsilon_{1},\ldots,\varepsilon_{p} such that εp2=1\varepsilon_{p}^{2}=-1. Conventionally, a 1-1-multigraded bundle is defined to be an ungraded bundle and a 0-multigraded bundle is simply a graded bundle. We say that a differential operator DD on SS is graded if it sends smooth sections of S+S^{+} to smooth sections of SS^{-} and vice-versa, and it is pp-multigraded if additionally it commutes with all of the multigrading operators.

Any pp-multigraded Dirac-type operator DD on a complete Riemannian manifold MM determines a class [D][D] in the degree pp K-homology group Kp(M)K_{p}(M); see chapter 10 of [3] for the details. The standard example of a pp-multigraded Dirac-type operator is the spinor Dirac operator on a pp-dimensional spin- or spinc-manifold. If MM is a compact manifold then any graded Dirac-type operator DD on MM is Fredholm, and the Fredholm index of DD depends only on its K-homology class. Indeed, there is a group homomorphism

K0(M)K_{0}(M)\to\mathbb{Z}

which sends the K-homology class of DD to its Fredholm index. In fact, this map is a special case of the coarse assembly map: since MM is compact it is coarsely equivalent to a point, and the coarse algebra of a point is simply the C*-algebra of compact operators. Thus the coarse assembly map in degree 0 is simply a map K0(M)K0(𝕂)K_{0}(M)\to K_{0}(\mathbb{K})\cong\mathbb{Z}, and this turns out to be the index map described above.

The coarse assembly map can be used to construct more general index maps. Let MM once again be a compact Riemannian manifold, let GG be its fundamental group, and let M~\widetilde{M} be its universal cover. Since MM is compact, M~\widetilde{M} is coarsely equivalent to GG (regarded as a metric space via any choice of word metric) and one can show that CG(M~)C_{G}^{*}(\widetilde{M}) is isomorphic to the reduced group C*-algebra Cr(G)C_{r}^{*}(G) of GG (see [3], chapter 12). Thus the coarse index map in this case is a map

Kp(M)Kp(Cr(G))K_{p}(M)\to K_{p}(C_{r}^{*}(G))

often called the higher index map or sometimes the equivariant index map.

4.2. The Partitioned Manifold Index Theorem

We are now ready to formulate and prove the partitioned manifold index theorem.

Definition 4.2.

Let MM be a smooth manifold and let NN be a submanifold of codimension 11. MM is partitioned by NN if MM is the union of two submanifolds M+M^{+} and MM^{-} with common boundary NN.

Assume now that MM is a complete Riemannian manifold (in particular a proper metric space) and that NN is compact. Let GG be a countable discrete group and let M~M\widetilde{M}\to M be a locally isometric GG-cover of MM. Define a map MM\to\mathbb{R} by the formula xdist(x,N)x\mapsto\text{dist}(x,N); this is a coarse map since MM is a length metric space and it lifts to a GG-invariant coarse map M~×G\widetilde{M}\to\mathbb{R}\times G. Thus it induces a homomorphism Kp(CG(M))Kp(CG(×G))K_{p}(C_{G}^{*}(M))\to K_{p}(C_{G}^{*}(\mathbb{R}\times G)). According to standard results in coarse K-theory ([3], chapter 6),

Kp(CG([0,)×G))=0K_{p}(C_{G}^{*}([0,\infty)\times G))=0

and thus the boundary map in the Mayer-Vietoris sequence for coarse K-theory associated to the decomposition ×G=(,0]×G[0,)×G\mathbb{R}\times G=(-\infty,0]\times G\cup[0,\infty)\times G is an isomorphism:

Kp(CG(×G))Kp1(Cr(G))K_{p}(C_{G}^{*}(\mathbb{R}\times G))\cong K_{p-1}(C_{r}^{*}(G))

This allows us to construct an index map associated to a partitioned manifold.

Definition 4.3.

Let MM be a complete Riemannian manifold partitioned by a compact hypersurface NN and let M~\widetilde{M} be a GG-cover of MM where GG is a countable discrete group. The partitioned index map is the composition

IndM,NG:Kp(M)Kp(CG(×M~))Kp1(Cr(G))\text{Ind}_{M,N}^{G}\colon K_{p}(M)\to K_{p}(C_{G}^{*}(\mathbb{R}\times\widetilde{M}))\cong K_{p-1}(C_{r}^{*}(G))

The partitioned manifold index theorem relates the partitioned index of a Dirac-type operator DD on MM to the ordinary (equivariant) index of its restriction to NN. For this to be possible we must make certain assumptions about the local structure of DD near NN:

Definition 4.4.

Let MM be a smooth manifold partitioned by a hypersurface NN, let SMMS_{M}\to M be a smooth pp-multigraded vector bundle over MM, and let SNNS_{N}\to N be a smooth (p1)(p-1)-multigraded vector bundle over NN. Let DMD_{M} and DND_{N} be pp- and (p1)(p-1)-multigraded differential operators acting on smooth sections of SMS_{M} and SNS_{N}, respectively. Say that DMD_{M} is partitioned by DND_{N} if there is a collaring neighborhood U(1,1)×NU\cong(-1,1)\times N of NN in MM with the following properties:

  • SM|US(1,1)^SNS_{M}|_{U}\cong S_{(-1,1)}\hat{\otimes}S_{N} where S(1,1)S_{(-1,1)} is the standard 11-multigraded complex spinor bundle over (1,1)(-1,1).

  • DM=D(1,1)^1+1^DND_{M}=D_{(-1,1)}\hat{\otimes}1+1\hat{\otimes}D_{N} where D(1,1)D_{(-1,1)} is the complex spinor Dirac operator on S(1,1)S_{(-1,1)}.

We are now ready to state the main theorem:

Theorem 4.5 (The Partitioned Manifold Index Theorem).

Let MM be a complete Riemannian manifold and let M~\widetilde{M} be a locally isometric GG-cover of MM where GG is a countable discrete group. Suppose MM is partitioned by a compact hypersurface NN and M~\widetilde{M} is partitioned by the lift N~\widetilde{N} of NN. If DMD_{M} is a pp-multigraded Dirac-type operator on MM which is partitioned by a (p1)(p-1)-multigraded Dirac-type operator DND_{N} on NN then

IndM,NG[DM]=IndNG[DN]\text{Ind}_{M,N}^{G}[D_{M}]=\text{Ind}_{N}^{G}[D_{N}]

in Kp1(Cr(G))K_{p-1}(C_{r}^{*}(G)).

Earlier proofs of this theorem, such as in [8] or [2], deal only with the case where GG is the trivial group. Additionally, earlier arguments gave little insight into the where NN is non-compact; the proof here uses the compactness hypothesis only to relate the coarse assembly map to the more familiar index map IndNG\text{Ind}_{N}^{G}. There are a variety of tools for calculating this; if GG is trivial then IndN\text{Ind}_{N} can be calculated explicitly in terms of characteristic classes on NN using the Atiyah-Singer index theorem, and for nontrivial GG IndNG\text{Ind}_{N}^{G} can be calculated in many examples using representation theory and equivariant K-theory.

The strategy of the proof is to fit the relevant index maps into a commutative diagram with the boundary map in the Mayer-Vietoris sequence for K-homology:

Kp(M)\textstyle{K_{p}(M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}IndM,NG\scriptstyle{\text{Ind}_{M,N}^{G}}MV\scriptstyle{\partial_{MV}}Kp1(Cr(G))\textstyle{K_{p-1}(C_{r}^{*}(G))}Kp1(N)\textstyle{K_{p-1}(N)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}IndNG\scriptstyle{\text{Ind}_{N}^{G}}

The result would then follow if we showed that MV\partial_{MV} sends the K-homology class of DMD_{M} to the K-homology class of DND_{N}. It is here that the technical assumptions on DMD_{M} in the statement of Theorem 4.5 play a crucial role. This is also where our calculations with Mayer-Vietoris boundary maps enter into the proof.

Lemma 4.6.

In the setting of Theorem 4.5 we have

MV[DM]=[DN]\partial_{MV}[D_{M}]=[D_{N}]

where [DM][D_{M}] is the class in Kp(M)K_{p}(M) determined by DMD_{M}, [DN][D_{N}] is the class in Kp1(N)K_{p-1}(N) determined by DND_{N}, and

MV:Kp(M)Kp1(M+M)=Kp1(N)\partial_{MV}\colon K_{p}(M)\to K_{p-1}(M^{+}\cap M^{-})=K_{p-1}(N)

is the boundary map in the Mayer-Vietoris sequence in K-homology associated to the topologically excisive decomposition M=M+MM=M^{+}\cup M^{-} determined by the partitioning of MM.

Proof.

Let φ+\varphi^{+} and φ\varphi^{-} denote the multiplication operators by the characteristic functions of M+M^{+} and MM^{-}, respectively. Observe that these operators are projections and (φ+,φ)(\varphi^{+},\varphi^{-}) defines a partition of unity for the decomposition 𝔇(M)=𝔇(M+M)+𝔇(MM)\mathfrak{D}^{*}(M)=\mathfrak{D}^{*}(M^{+}\subseteq M)+\mathfrak{D}^{*}(M^{-}\subseteq M). The K-homology class [DM]Kp(M)=K1p(𝔇(M))[D_{M}]\in K_{p}(M)=K_{1-p}(\mathfrak{D}^{*}(M)) is represented by an operator FF in a matrix algebra over 𝔇(M)\mathfrak{D}^{*}(M) (FF is either a projection or a unitary depending on the multigrading structure of DMD_{M}), and MV\partial_{MV} sends [F][F] to the image of [φ+F+φ]K1p(𝔇(M)/𝔇(NM))[\varphi^{+}F+\varphi^{-}]\in K_{1-p}(\mathfrak{D}^{*}(M)/\mathfrak{D}^{*}(N\subseteq M)) under the boundary map

(4.1) K1p(𝔇(M)/𝔇(NM))Kp(𝔇(NM))Kp1(N)K_{1-p}(\mathfrak{D}^{*}(M)/\mathfrak{D}^{*}(N\subseteq M))\to K_{-p}(\mathfrak{D}^{*}(N\subseteq M))\cong K_{p-1}(N)

According to the excision theorem in K-homology ([3], chapter 5),

K1p(𝔇(M)/𝔇(NM))Kp(MN)Kp(M+N)K_{1-p}(\mathfrak{D}^{*}(M)/\mathfrak{D}^{*}(N\subseteq M))\cong K_{p}(M^{-}-N)\oplus K_{p}(M^{+}-N)

so MV\partial_{MV} sends [DM][D_{M}] to [DM]\partial[D_{M}] where

:Kp(M+N)Kp1(N)\partial\colon K_{p}(M^{+}-N)\to K_{p-1}(N)

is the boundary map in K-homology. Because of the local structure of DMD_{M} near NN, standard results in analytic K-homology ([3], chapter 11) imply that MV[DM]=[DN]\partial_{MV}[D_{M}]=[D_{N}] (this is essentially an application of the slogan, ”the boundary of Dirac is Dirac”). ∎

The proof of the partitioned manifold index theorem now follows immediately:

Proof of Theorem 4.5.

Using the Mayer-Vietoris and analytic surgery diagram, there is a commuting diagram:

Kp(M)\textstyle{K_{p}(M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}MV\scriptstyle{\partial_{MV}}IndM,NG\scriptstyle{\text{Ind}_{M,N}^{G}}Kp(CG(M~×G))\textstyle{K_{p}(C_{G}^{*}(\widetilde{M}\times G))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}MV\scriptstyle{\partial_{MV}}Kp1(N)\textstyle{K_{p-1}(N)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}IndNG\scriptstyle{\text{Ind}_{N}^{G}}Kp1(Cr(G))\textstyle{K_{p-1}(C_{r}^{*}(G))}

Commutativity of this diagram together with Lemma 4.6 completes the proof. ∎

4.3. Generalizations and Applications

The main idea of Theorem 4.5 is to compute the index of an operator on a non-compact manifold MM by localizing the calculation to a compact partitioning hypersurface. This is possible because the partitioning structure provides a mechanism for relating the uniform geometry of MM to the uniform geometry of \mathbb{R}. This in turn suggests that a similar result can be proved for a manifold whose uniform geometry is related to that of k\mathbb{R}^{k}.

Definition 4.7.

Let MM be a smooth manifold and let NN be a submanifold of codimension kk. A kk-partitioning map for the pair (M,N)(M,N) is a coarse submersion F:MkF\colon M\to\mathbb{R}^{k} such that N=F1(0)N=F^{-1}(0). Say that MM is kk-partitioned by NN if there exists a kk-partitioning map for (M,N)(M,N).

We shall consider operators on MM which have a favorable local structure near NN, just as before:

Definition 4.8.

Let MM be a smooth manifold and suppose NN is a submanifold of MM of codimension kk. Let DMD_{M} be a pp-multigraded differential operator acting on a smooth pp-multigraded vector bundle SMMS_{M}\to M, pkp\geq k, and let DND_{N} be a (pk)(p-k)-multigraded differential operator acting on a smooth (pk)(p-k)-multigraded vector bundle SNNS_{N}\to N. Say that DMD_{M} is kk-partitioned by DND_{N} if there is a collaring neighborhood U(1,1)k×NU\cong(-1,1)^{k}\times N of NN in MM with the following properties:

  • SM|US(1,1)k^SNS_{M}|_{U}\cong S_{(-1,1)^{k}}\hat{\otimes}S_{N} where S(1,1)kS_{(-1,1)^{k}} is the complex spinor bundle on the cube (1,1)k(-1,1)^{k}

  • DM=D(1,1)k^1+1^DND_{M}=D_{(-1,1)^{k}}\hat{\otimes}1+1\hat{\otimes}D_{N} where D(1,1)kD_{(-1,1)^{k}} is the complex spinor Dirac operator

Suppose MM is kk-partitioned by a compact hypersurface NN and M~\widetilde{M} is a locally isometric GG-cover of MM where GG is a countable discrete group. Suppose further that the kk-partitioning map F:MkF\colon M\to\mathbb{R}^{k} lifts to a kk-partitioning map F~:M~k\widetilde{F}\colon\widetilde{M}\to\mathbb{R}^{k} for the pair (M~,N~)(\widetilde{M},\widetilde{N}). Then F~\widetilde{F} induces a map

F~:Kp(CG(M~))Kp(CG(k×G))\widetilde{F}_{*}\colon K_{p}(C_{G}^{*}(\widetilde{M}))\to K_{p}(C_{G}^{*}(\mathbb{R}^{k}\times G))

By induction on kk there is an isomorphism Kp(CG(k×G))Kpk(Cr(G))K_{p}(C_{G}^{*}(\mathbb{R}^{k}\times G))\cong K_{p-k}(C_{r}^{*}(G)) given by iterated Mayer-Vietoris boundary maps. Thus we have a kk-partitioned index map:

IndM,NG:Kp(M)Kp(CG(M~))Kpk(Cr(G))\text{Ind}_{M,N}^{G}\colon K_{p}(M)\to K_{p}(C_{G}^{*}(\widetilde{M}))\cong K_{p-k}(C_{r}^{*}(G))

analogous to the partitioned index map defined above. Our main result about kk-partitioned manifolds computes this index map.

Proposition 4.9.

Let MM be a complete Riemannian manifold and let M~\widetilde{M} be a locally isometric GG-cover of MM where GG is a countable discrete group. Let NN be a submanifold of MM which lifts to a submanifold N~\widetilde{N} of M~\widetilde{M}, and suppose there is a kk-partitioning map F:MkF\colon M\to\mathbb{R}^{k} for the pair (M,N)(M,N) which lifts to a kk-partitioning map F~:M~k\widetilde{F}\colon\widetilde{M}\to\mathbb{R}^{k} for the pair (M~,N~)(\widetilde{M},\widetilde{N}). If DMD_{M} is a pp-multigraded Dirac-type operator on MM, pkp\geq k, which is kk-partitioned by a (pk)(p-k)-multigraded Dirac-type operator DND_{N} on NN then

IndM,NG[DM]=IndNG[DN]\text{Ind}_{M,N}^{G}[D_{M}]=\text{Ind}_{N}^{G}[D_{N}]

in Kpk(Cr(G))K_{p-k}(C_{r}^{*}(G)).

Proof.

We use induction on kk; the base case is simply Theorem 4.5, so assume k2k\geq 2. Since FF is a submersion the sets M+=F1(k1×0)M^{+}=F^{-1}(\mathbb{R}^{k-1}\times\mathbb{R}^{\geq 0}) and M=F1(k1×0)M^{-}=F^{-1}(\mathbb{R}^{k-1}\times\mathbb{R}^{\leq 0}) are submanifolds with boundary which partition MM, and the partitioning hypersurface N=M+MN^{\prime}=M^{+}\cap M^{-} is (k1)(k-1)-partitioned by NN. Moreover DMD_{M} is partitioned by the (k1)(k-1)-multigraded operator DN=D(1,1)k1^1+1^DND_{N^{\prime}}=D_{(-1,1)^{k-1}}\hat{\otimes}1+1\hat{\otimes}D_{N} and DND_{N^{\prime}} is (k1)(k-1)-partitioned by DND_{N}. By the induction hypothesis it suffices to show that IndM,NG[DM]=IndN,N[DN]\text{Ind}_{M,N}^{G}[D_{M}]=\text{Ind}_{N^{\prime},N}[D_{N^{\prime}}] in Kpk(Cr(G))K_{p-k}(C_{r}^{*}(G)). As in the proof of Theorem 4.5 there is a commutative diagram

Kp(M)\textstyle{K_{p}(M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}IndM,NG\scriptstyle{\text{Ind}_{M,N}^{G}}MV\scriptstyle{\partial_{MV}}Kpk(Cr(G))\textstyle{K_{p-k}(C_{r}^{*}(G))}Kp1(N)\textstyle{K_{p-1}(N^{\prime})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}IndN,NG\scriptstyle{\text{Ind}_{N^{\prime},N}^{G}}

so the result follows from Lemma 4.6. ∎

We conclude by providing an application of Proposition 4.9 to the theory of positive scalar curvature invariants. A consequence of the classical Lichnerowicz formula is that if MM is a complete Riemannian spin manifold whose scalar curvature function is bounded below by a positive constant and DD is the spinor Dirac operator on MM then the K-homology class of DD lifts to the analytic structure group ([3], chapter 12). According to the analytic surgery exact sequence

Kp+1(C(M))Sp(M)Kp(M)Kp(C(M))\to K_{p+1}(C^{*}(M))\to S_{p}(M)\to K_{p}(M)\to K_{p}(C^{*}(M))\to

this implies that the coarse assembly map sends [D][D] to 0 in Kp(C(M))K_{p}(C^{*}(M)). This observation together with Proposition 4.9 yield a new proof of a celebrated theorem of Gromov and Lawson ([1]).

Theorem 4.10 (Gromov and Lawson).

Let MM be a compact manifold of dimension nn which admits a Riemannian metric of non-positive sectional curvature. Then MM has no metric of positive scalar curvature.

Proof.

Let M~\widetilde{M} be the universal cover of MM equipped with a Riemannian metric of non-positive sectional curvature (lifted from a metric gg on MM). M~\widetilde{M} is simply connected, so according to the Cartan-Hadamard theorem the exponential map expp:TpMM\exp_{p}\colon T_{p}M\to M is a diffeomorphism for any pp. In fact we can say more: the exponential map is expansive in the sense that d(expp(v1),expp(v2))v1v2d(\exp_{p}(v_{1}),\exp_{p}(v_{2}))\geq\left\|v_{1}-v_{2}\right\| for any v1,v2TpMv_{1},v_{2}\in T_{p}M, so the inverse map log:M~TpM\log\colon\widetilde{M}\to T_{p}M is a coarse diffeomorphism.

Now, let gg^{\prime} be any Riemannian metric on MM. Any two norms on a finite dimensional vector space are equivalent, so there is a constant CC such that 1CvvC\frac{1}{C}\left\|v\right\|^{\prime}\leq\left\|v\right\|\leq C for every vTpMv\in T_{p}M. Integrating, it follows that 1Cd(x,y)d(x,y)Cd(x,y)\frac{1}{C}d^{\prime}(x,y)\leq d(x,y)\leq Cd^{\prime}(x,y) for every x,yMx,y\in M, so it follows that the map log:M~TpM\log\colon\widetilde{M}\to T_{p}M defined using the metric gg is still a coarse map with respect to the metric gg^{\prime} (lifted to M~\widetilde{M}). Thus the 0-dimensional submanifold {p}\{p\} of M~\widetilde{M} is nn-partitioned by the coarse diffeomorphism log:M~TpMn\log\colon\widetilde{M}\to T_{p}M\cong\mathbb{R}^{n}.

Clearly M~\widetilde{M} is contractible, so let SM~S_{\widetilde{M}} denote the trivial spinor bundle n×M~\mathbb{R}_{n}\times\widetilde{M} and let DM~D_{\widetilde{M}} denote the spinor Dirac operator on SM~S_{\widetilde{M}} (defined using the Riemannian metric gg^{\prime}). Consider the trivial bundle Sp=×{p}{p}S_{p}=\mathbb{C}\times\{p\}\to\{p\} and let DpD_{p} be the zero map on SpS_{p}; DM~D_{\widetilde{M}} is kk-partitioned by DpD_{p} because SM~n^SpS_{\widetilde{M}}\cong\mathbb{R}_{n}\hat{\otimes}S_{p} and DM~=Dn^1+1^DpD_{\widetilde{M}}=D_{\mathbb{R}^{n}}\hat{\otimes}1+1\hat{\otimes}D_{p}. By the index theorem for kk-partitioned manifolds, we have

IndM~,{p}(DM~)=Ind{p}(Dp)=1\text{Ind}_{\widetilde{M},\{p\}}(D_{\widetilde{M}})=\text{Ind}_{\{p\}}(D_{p})=1

in \mathbb{Z}. But IndM~,{p}\text{Ind}_{\widetilde{M},\{p\}} factors through the coarse index map

Kp(M~)Kp(C(M~))K_{p}(\widetilde{M})\to K_{p}(C^{*}(\widetilde{M}))

and the image under this map of DM~D_{\widetilde{M}} would be 0 if the scalar curvature function on MM associated to gg^{\prime} were positive. Thus MM can have no metric of positive scalar curvature. ∎

References

  • [1] Mikhael Gromov and Blaine Lawson. Positive scalar curvature and the dirac operator on complete riemannian manifolds. Inst. Hautes Etudes Sci. Publ. Math., (58):83–196, 1984.
  • [2] Nigel Higson. A note on the cobordism invariance of the index. Topology, 1991.
  • [3] Nigel Higson and John Roe. Analytic K-Homology. Oxford University Press, 2000.
  • [4] Nigel Higson and John Roe. Mapping surgery to analysis I: Analytic signatures. K-Theory, 2005.
  • [5] Nigel Higson and John Roe. Mapping surgery to analysis II: Geometric signatures. K-Theory, 2005.
  • [6] Nigel Higson and John Roe. Mapping surgery to analysis III: Exact sequences. K-Theory, 2005.
  • [7] John Roe Nigel Higson and Guo Liang Yu. A coarse mayer-vietoris principle. Math. Proc. Cambridge Philos. Soc., 1993.
  • [8] John Roe. Partitioning noncompact manifolds and the dual toeplitz problem. In Operator Algebras and Applications, Vol 1. Cambridge University Press, 1988.
  • [9] John Roe and Paul Siegel. Sheaf theory and paschke duality. To Appear, 2012.
  • [10] Paul Siegel. Homological Calculations with Analytic Structure Groups. PhD thesis, Penn State, 2012.