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

Band width and the Rosenberg index

Yosuke Kubota Department of Mathematical Sciences, Shinshu University
3-1-1 Asahi, Matsumoto, Nagano, 390-8621, Japan
and
RIKEN iTHEMS
2-1 Hirosawa, Wako, Saitama, 351-0198, Japan
[email protected]
Abstract.

A Riemannian manifold is said to have infinite 𝒦​π’ͺ\mathcal{KO}-width if it admits an isometric immersion of an arbitrarily wide Riemannian band whose inward boundary has non-trivial higher index. In this paper we prove that if a closed spin manifold has inifinite 𝒦​π’ͺ\mathcal{KO}-width, then its Rosenberg index does not vanish. This gives a positive answer to a conjecture by R.Β Zeidler. We also prove its β€˜multi-dimensional’ generalization; if a closed spin manifold admit an isometric immersion of an arbitrarily wide cube-like domain whose lowest dimensional corner has non-trivial higher index, then the Rosenberg index of MM does not vanish.

1. Introduction

The existence of a positive scalar curvature (psc) metric on a given manifold has been a fundamental problem in high-dimensional differential topology. An effective approach is the Dirac operator method, in which the SchrΓΆdinger–Lichnerowicz theorem reduces the problem to the invertibility of the Dirac operator. When the manifold is not compact, the invertibility of the Dirac operator is obstructed by the higher index, a generalization of the Fredholm index defined by using C*-algebra K-theory and coarse geometry [roeLecturesCoarseGeometry2003, willett_yu_2020]. When one consider the universal covering of a closed manifold, the higher index of the Dirac operator is called the Rosenberg index [rosenbergAstAlgebrasPositive1983, rosenbergAstAlgebrasPositive1986, rosenbergAstAlgebrasPositive1986a] and is known to be a powerful obstruction to a psc metric. Indeed, the Rosenberg–Stolz theorem [rosenbergStableVersionGromovLawson1995, stolzManifoldsPositiveScalar2002] states that the Rosenberg index is a complete obstruction to positive scalar curvature in a stable sense, under the assumption of the Baum–Connes injectivity. More precisely, the vanishing of the Rosenberg index is equivalent to the existence of a psc metric after taking the direct product with sufficiently many copies of the Bott manifold (an 88-dimensional closed spin manifold with Sgn​(B)=0\mathrm{Sgn}(B)=0 and A^​(B)=1\hat{A}(B)=1). On the other hand, Schick [schickCounterexampleUnstableGromovLawsonRosenberg1998] constructed a closed spin manifold in dimensions 55, 66, 77 which does not admit any psc metric but its Rosenberg index vanishes, by using the Schoen–Yau minimal surface method [schoenStructureManifoldsPositive1979]. This leads us to explore a psc obstruction beyond the Rosenberg index. A guideline is Schick’s meta-conjectureΒ [schickTopologyPositiveScalar2014]*Conjecture 1.5, stating that any topological obstruction to positive scalar curvature coming from the Dirac operator is dominated by the Rosenberg index. For example, the psc obstructions given by the Rosenberg index of certain submanifolds of codimension 1 [zeidlerIndexObstructionPositive2017, kubotaRelativeMishchenkoFomenkoHigher2020] and codimension 22 [hankeCodimensionTwoIndex2015, kubotaRelativeMishchenkoFomenkoHigher2019, kubotaGromovLawsonCodimensionObstruction2020, kubotaCodimensionTransferHigher2021] provide evidences to this conjecture.

Recently, a series of papers by Gromov [gromovDozenProblemsQuestions2018, gromovMetricInequalitiesScalar2018, gromovFourLecturesScalar2019] shed new lights on this problem. One of the remarkable ideas proposed in these papers is the notion of a band and its width. A (proper) compact Riemannian band VV is a compact Riemannian manifold with inward and outward boundaries βˆ‚Β±V\partial_{\pm}V. The distance of βˆ‚+V\partial_{+}V and βˆ‚βˆ’V\partial_{-}V is called its width. Gromov proved that, if a compact Riemannian band VV is endowed with a psc metric but βˆ‚+V\partial_{+}V does not admit any psc metric due to the minimal surface method, then the width of VV is bounded by a constant depending on the infimum of the scalar curvature and the dimension. Following this line, in [zeidlerBandWidthEstimates2020, zeidlerWidthLargenessIndex2020, cecchiniLongNeckPrinciple2020], Zeidler and Cecchini proved the same band width inequality when the inward boundary βˆ‚+V~\partial_{+}\widetilde{V} of the universal covering V~\widetilde{V} has non-trivial higher index. Furthermore, another approach to this inequality based on the quantitative K-theory is developed by Guo–Xie–YuΒ [guoQuantitativeKtheoryPositive2020].

This band width inequality has a qualitative application to the existence of a psc. It is considered by Gromov [gromovMetricInequalitiesScalar2018]*Section 3,4 and Zeidler [zeidlerBandWidthEstimates2020]*Section 4, the latter of which is the main subject of this paper. The following definition of the notion of 𝒦​π’ͺ\mathcal{KO}-width looks a little different, but equivalent, to the original definition by Zeidler.

Definition 1.1 ([zeidlerBandWidthEstimates2020]*Definition 4.3).

A compact Riemannian band is said to be in the class 𝒦​π’ͺ\mathcal{KO} if it is equipped with a spin structure and the index of the higher index IndΟ€1​(V)⁑(DΜΈβˆ‚+V~)\operatorname{\mathrm{Ind}}_{\pi_{1}(V)}(\not{D}_{\partial_{+}\widetilde{V}}) does not vanish, where βˆ‚+V~\partial_{+}\widetilde{V} denotes the inward boundary of the universal covering of VV. For a closed manifold MM, its 𝒦​π’ͺ\mathcal{KO}-width width𝒦​π’ͺ​(M,g)\mathrm{width}_{\mathcal{KO}}(M,g) is the supremum of the width of bands in the class 𝒦​π’ͺ\mathcal{KO} which is isometrically immersed to MM.

Note that the infiniteness of the 𝒦​π’ͺ\mathcal{KO}-width depends only on the diffeomorphism class of MM, i.e., is independent of the metric on it. The band width inequality implies that a closed manifold with infinite 𝒦​π’ͺ\mathcal{KO}-width does not admit a psc metric. Since the infiniteness of the 𝒦​π’ͺ\mathcal{KO}-width is stable under the direct product with the Bott manifold, the Rosenberg–Stolz theorem shows the non-vanishiing of the Rosenberg index of MM, if we assume that Ο€1​(M)\pi_{1}(M) satisfies the Baum–Connes injectivity. Zeidler conjectured that this non-vanishing still holds without assuming the Baum–Connes injectivity [zeidlerBandWidthEstimates2020]*Conjecture 4.12, following the line of Schick’s meta-conjecture. The aim of this paper is to give a positive answer to this conjecture.

In this paper we work in a little more general setting. We need not assume that the target manifold MM is closed. Instead, we assume that the universal covering M~\widetilde{M} of MM has a well-behaved β€˜uniform’ topology. For x∈Xx\in X and R>0R>0, let BR​(x)B_{R}(x) denote the ball with the center xx and the radius RR.

Definition 1.2.

A metric space XX is said to be uniformly 11-connected if there is an increasing function Ο†:ℝ>0→ℝ>0\varphi\colon\mathbb{R}_{>0}\to\mathbb{R}_{>0} with φ​(t)β†’βˆž\varphi(t)\to\infty as tβ†’βˆžt\to\infty such that any loop in BR​(x)B_{R}(x) is trivial in Bφ​(R)​(x)B_{\varphi(R)}(x).

Here we extend the notion of 𝒦​π’ͺ\mathcal{KO}-width to complete Riemannian manifolds. Then its infiniteness depends only on the diffeomorphism class of MM and the coarse equivalence class of the metric. Now we state the first main theorem of this paper.

Theorem 1.3.

Let (M,g)(M,g) be a complete Riemannian spin manifold whose universal covering M~\widetilde{M} is uniformly 11-connected. If MM has infinite 𝒦​π’ͺ\mathcal{KO}-width, then the maximal equivariant coarse index IndΓ⁑(DΜΈM~)\operatorname{\mathrm{Ind}}_{\Gamma}(\not{D}_{\widetilde{M}}) of the Dirac operator DΜΈM~\not{D}_{\widetilde{M}} on M~\widetilde{M} does not vanish.

Corollary 1.4.

Let MM be a closed spin manifold. If MM has infinite 𝒦​π’ͺ\mathcal{KO}-width, then the Rosenberg index αΓ​(M)\alpha_{\Gamma}(M) does not vanish.

We remark that this corollary reproves [kubotaGromovLawsonCodimensionObstruction2020]*Theorem 1.1 as a special case, as is pointed out by Zeidler in [zeidlerBandWidthEstimates2020]*Example 4.9.

There are two ingredients of the proof. One is the asymptotic method in C*-algebra K-theory. The asymptotic C*-algebras, typically the quotient of the direct product of a sequence of C*-algebras by the direct sum, has been exploited in many researches of higher index theory such as Hanke–SchickΒ [hankeEnlargeabilityIndexTheory2006, hankeEnlargeabilityIndexTheory2007], Gong–Wang–YuΒ [gongGeometrizationStrongNovikov2008], and so on. In this paper, the C*-algebra of this kind leads us to a qualitative treatment of the higher index of a sequence {Vn}nβˆˆβ„•\{V_{n}\}_{n\in\mathbb{N}} of Riemannian bands getting wider as nβ†’βˆžn\to\infty, which suits our purpose (although the quantitative estimate is the highlight of this new research direction after Gromov).

Another ingredient is an estimate of the (relative) systole of Riemannian bands immersed to MM. What makes the problem seem difficult is that the immersion Vβ†’MV\to M does not induce the injection of fundamental groups in general. However, we show in LemmaΒ 2.1 that the length of a non-trivial loop in VV, which is apart from the boundary and trivial in MM, is bounded below by a constant depending on the width of VV. This enables us to lift an operator on M~\widetilde{M} to the universal covering V~\widetilde{V} of VV β€˜modulo the boundary’, which is a variant of the lifting lemma developed in [kubotaCodimensionTransferHigher2021].

We also discuss a generalization of TheoremΒ 1.3, in which we consider the multiwidth of cube-like domains immersed to MM instead of the width of bands. We say that a β–‘m\square^{m}-domain is a Riemannian manifold with corner VV equipped with a well-behaved corner-preserving smooth map f:Vβ†’[βˆ’1,1]mf\colon V\to[-1,1]^{m} (more precisely, see DefinitionΒ 3.1). We write βˆ‚j,Β±V\partial_{j,\pm}V for the inverse image fβˆ’1​({xj=Β±1})f^{-1}(\{x_{j}=\pm 1\}). The multiwidth of VV is defined by the minimum of the distances of βˆ‚j,+V\partial_{j,+}V and βˆ‚j,βˆ’V\partial_{j,-}V. In recent researches, a generalization of the band width inequality to such domains has been considered; if VV is a β–‘m\square^{m}-domain equipped with a psc metric but its lowest dimensional corner Vβ‹”V_{\pitchfork} (more precisely, a transverse intersection of mm hypersurfaces each of which separates βˆ‚j,+V\partial_{j,+}V with βˆ‚j,βˆ’V\partial_{j,-}V) does not admit a psc metric due to the Dirac operator or minimal surface methods, then its multiwidth is bounded by a constant depending on the infimum of the scalar curvature and the dimension. This inequality, called the β–‘nβˆ’m\square^{n-m}-theorem or the β–‘nβˆ’m\square^{n-m}-inequality, is proposed and proved for dimV≀8\dim V\leq 8 by Gromov [gromovFourLecturesScalar2019]*p.260, and generalized to manifolds with all dimensions by Wang–Xie–Yu and Xie [wangProofGromovCube2021, xieQuantitativeRelativeIndex2021] in the case that the psc metric on Vβ‹”V_{\pitchfork} is obstructed by its Rosenberg index. A qualitative consequence of this inequality is that if a closed spin manifold MM has infinite 𝒦​π’ͺm\mathcal{KO}_{m}-multiwidth (DefinitionΒ 3.3), then MM does not admit any psc metric. The second main theorem of this paper is to dominate this obstruction to positive scalar curvature by the Rosenberg index.

Theorem 1.5.

Let (M,g)(M,g) be a complete Riemannian spin manifold whose universal covering M~\widetilde{M} is uniformly 11-connected. If MM has infinite 𝒦​π’ͺm\mathcal{KO}_{m}-multiwidth (DefinitionΒ 3.3), then the maximal coarse index IndΓ⁑(DΜΈM~)∈KOd⁑(Cβˆ—β€‹(M~)Ξ“)\operatorname{\mathrm{Ind}}_{\Gamma}(\not{D}_{\widetilde{M}})\in\operatorname{\mathrm{KO}}_{d}(C^{*}(\widetilde{M})^{\Gamma}) of the Dirac operator on M~\widetilde{M} does not vanish.

Corollary 1.6.

Let (M,g)(M,g) be a closed Riemannian spin manifold. If MM has infinite 𝒦​π’ͺm\mathcal{KO}_{m}-width, then the Rosenberg index αΓ​(M)\alpha_{\Gamma}(M) does not vanish.

Acknowledgement

The author would like to thank Shinichiroh Matsuo for a helpful comment. This work was supported by RIKEN iTHEMS and JSPS KAKENHI Grant Numbers 19K14544, JPMJCR19T2, 17H06461.

2. Proof of TheoremΒ 1.3

2.1. Systole of Riemannian bands immersed to a uniformly 11-connected manifold

The first step for the proof is an observation on the systole, the minimum of the length of homotopically non-trivial loop, of a subspace of a uniformly 11-connected manifold.

Let MM be a dd-dimensional complete Riemannian spin manifold with infinite 𝒦​π’ͺ\mathcal{KO}-width. For each nβˆˆβ„•n\in\mathbb{N}, pick a compact Riemannian band VnV_{n} in the class 𝒦​π’ͺ\mathcal{KO} whose width is greater than nn, and an isometric immersion fn:Vnβ†’Mf_{n}\colon V_{n}\to M. Let Ο€n:=Ο€1​(Vn)\pi_{n}:=\pi_{1}(V_{n}) and let Ξ“nβŠ‚Ξ“\Gamma_{n}\subset\Gamma denote the image of Ο€n\pi_{n} under the inclusion Vnβ†’MV_{n}\to M. Then the map fnf_{n} lifts to a codimension 0 immersion f~n:V~nβ†’M~\tilde{f}_{n}\colon\widetilde{V}_{n}\to\widetilde{M}, where V~n\widetilde{V}_{n} is the universal covering of VnV_{n}. In short, we write 𝐕\mathbf{V} and 𝐕~\widetilde{\mathbf{V}} for the disjoint union ⨆nβˆˆβ„•Vn\bigsqcup_{n\in\mathbb{N}}V_{n} and ⨆nβˆˆβ„•V~n\bigsqcup_{n\in\mathbb{N}}\widetilde{V}_{n} respectively. For R>0R>0, let V~n,R\widetilde{V}_{n,R} denote the open subset of V~n\widetilde{V}_{n} consisting of points x∈V~nx\in\widetilde{V}_{n} such that d​(x,βˆ‚V~)β‰₯Rd(x,\partial\widetilde{V})\geq R.

Let U~n\widetilde{U}_{n} and U~n,R\widetilde{U}_{n,R} denote the interior of f~​(V~n)\tilde{f}(\widetilde{V}_{n}) and f~​(V~n,R)\tilde{f}(\widetilde{V}_{n,R}) respectively, which are non-empty open subsets of M~\widetilde{M}. For an inclusion YβŠ‚XY\subset X of length spaces, we call the infimum of the length of closed loops in YY representing non-trivial homotopy class in XX the relative systole of YβŠ‚XY\subset X and write sys​(YβŠ‚X)\mathrm{sys}(Y\subset X). This corresponds to the systole of YY relative to Ο€1​(Y)β†’Ο€1​(X)\pi_{1}(Y)\to\pi_{1}(X) in the standard terminology of systolic geometry (cf.Β [katzSystolicGeometryTopology2007]*Definition 8.2.1).

Lemma 2.1.

Assume that M~\widetilde{M} is uniformly 11-connected with respect to a function Ο†:ℝ>0→ℝ>0\varphi\colon\mathbb{R}_{>0}\to\mathbb{R}_{>0} and let CR:=inf{C>0βˆ£Ο†β€‹(C)β‰₯R}C_{R}:=\inf\{C>0\mid\varphi(C)\geq R\}. Then we have sys​(U~n,RβŠ‚U~n)β‰₯2​CR\mathrm{sys}(\widetilde{U}_{n,R}\subset\widetilde{U}_{n})\geq 2C_{R}, where the relative systole is defined with respect to the Riemannian distance of M~\widetilde{M}.

Proof.

Let β„“:S1β†’U~n,R\ell\colon S^{1}\to\widetilde{U}_{n,R} be a closed loop with length​(β„“)≀2​CR\mathrm{length}(\ell)\leq 2C_{R}. Let x:=ℓ​(0)x:=\ell(0). Note that β„“\ell is contained in the open ball BCR​(x)B_{C_{R}}(x). By the assumption of uniform 11-connectedness of M~\widetilde{M}, this β„“\ell is null-homotopic in Bφ​(CR)​(x)βŠ‚BR​(x)B_{\varphi(C_{R})}(x)\subset B_{R}(x). Since x∈U~n,Rx\in\widetilde{U}_{n,R}, we have BR​(x)βŠ‚U~nB_{R}(x)\subset\widetilde{U}_{n}. Hence we obtain that [β„“]βˆˆΟ€1​(U~n)[\ell]\in\pi_{1}(\widetilde{U}_{n}) is trivial. This shows that sys​(U~n,RβŠ‚U~n)β‰₯2​CR\mathrm{sys}(\widetilde{U}_{n,R}\subset\widetilde{U}_{n})\geq 2C_{R}. ∎

The following lemma is standard in coarse geometry, rather known as the uniform contractibility of the universal covering of an aspherical manifold (see e.g.Β [roeLecturesCoarseGeometry2003]*Example 5.26).

Lemma 2.2.

Let (M,g)(M,g) be a closed Riemannian manifold. Then the universal covering M~\widetilde{M} is uniformly 11-connected.

Proof.

For R>0R>0 and x∈M~x\in\widetilde{M}, let Cx,RC_{x,R} denote the infimum of the real numbers Cβ‰₯RC\geq R such that the inclusion BR​(x)β†’BC​(x)B_{R}(x)\to B_{C}(x) induces the trivial map in Ο€1\pi_{1}-groups. We show that Cx,R<∞C_{x,R}<\infty for any xx and RR. Let r>0r>0 be less than the injectivity radius of MM. Then BR​(x)B_{R}(x) is covered by a finite number of open balls Br​(yi)B_{r}(y_{i}) for i=1,β‹―,ki=1,\cdots,k. The open subspace U:=⋃i=1kBr​(yi)βŠ‚M~U:=\bigcup_{i=1}^{k}B_{r}(y_{i})\subset\widetilde{M} is homotopy equivalent to its nerve, and hence has the homotopy type of a finite simplicial complex. In particular, its fundamental group is finitely generated, and hence there is L>0L>0 such that UβŠ‚BL​(x)U\subset B_{L}(x) and the induced map Ο€1​(U)β†’Ο€1​(BL​(x))\pi_{1}(U)\to\pi_{1}(B_{L}(x)) is trivial. This shows Cx,R≀L<∞C_{x,R}\leq L<\infty, as desired.

The assignment x↦Cx,Rx\mapsto C_{x,R} is Ξ“\Gamma-invariant and locally bounded (indeed, Cy,R≀Cx,R+Ξ΅C_{y,R}\leq C_{x,R+\varepsilon} for any y∈BΡ​(x)y\in B_{\varepsilon}(x)). Therefore, together with compactness of M=M~/Ξ“M=\widetilde{M}/\Gamma, we get φ​(R):=supx∈M~(Cx,R+1)<∞\varphi(R):=\sup_{x\in\widetilde{M}}(C_{x,R}+1)<\infty. This Ο†\varphi is the desired function since it is by definition increasing. ∎

2.2. Lifting finite propagation operators

Next, we construct a lift of an operator on M~\widetilde{M} to the non-compact Riemannian bands V~n\widetilde{V}_{n}, which forms a βˆ—\ast-homomorphism β€˜modulo boundary’. This technique is inherited from [kubotaCodimensionTransferHigher2021].

Let Tβˆˆπ”Ήβ€‹(L2​(M~))T\in\mathbb{B}(L^{2}(\widetilde{M})) which is locally Hilbert–Schmidt, i.e., T​fTf, f​TfT is in the Hilbert–Schmidt class for any f∈Cc​(M~)f\in C_{c}(\widetilde{M}). Such TT is represented by a kernel function T:M~Γ—M~β†’β„‚T\colon\widetilde{M}\times\widetilde{M}\to\mathbb{C} as

T​ξ​(x):=∫y∈M~T​(x,y)​ξ​(y)​𝑑volg​(y).T\xi(x):=\int_{y\in\widetilde{M}}T(x,y)\xi(y)d\mathrm{vol}_{g}(y).

The support of TT is defined as the support of its kernel function in M~Γ—M~\widetilde{M}\times\widetilde{M}, and Prop​(T):=sup{d​(x,y)∣(x,y)∈supp⁑T}\mathrm{Prop}(T):=\sup\{d(x,y)\mid(x,y)\in\operatorname{\mathrm{supp}}T\} is called the propagation of TT, where dd denotes the metric on M~\widetilde{M}. We define β„‚HS​[M~]Ξ“\mathbb{C}_{\mathrm{HS}}[\widetilde{M}]^{\Gamma} as the Real βˆ—\ast-algebra of Ξ“\Gamma-invariant locally Hilbert–Schmidt operators with finite propagation. Let Cβˆ—β€‹(M~)Ξ“C^{*}(\widetilde{M})^{\Gamma} denote the completion of β„‚HS​[M~]Ξ“\mathbb{C}_{\mathrm{HS}}[\widetilde{M}]^{\Gamma} with respect to the maximal norm satisfying the C*-condition β€–Tβˆ—β€‹Tβ€–=β€–Tβ€–2\|T^{*}T\|=\|T\|^{2} (which is well-defined by [gongGeometrizationStrongNovikov2008]*3.5). In the same way, we also define β„‚HS​[V~n]Ο€nβŠ‚π”Ήβ€‹(L2​(V~n))\mathbb{C}_{\mathrm{HS}}[\widetilde{V}_{n}]^{\pi_{n}}\subset\mathbb{B}(L^{2}(\widetilde{V}_{n})) and its completion Cβˆ—β€‹(V~n)Ο€nC^{*}(\widetilde{V}_{n})^{\pi_{n}}.

We define the Real βˆ—\ast-algebra β„‚HS​[𝐕~]𝝅\mathbb{C}_{\mathrm{HS}}[\widetilde{\mathbf{V}}]^{\boldsymbol{\pi}} consisting of sequences of locally compact operators Tnβˆˆπ”Ήβ€‹(L2​(V~n))T_{n}\in\mathbb{B}(L^{2}(\widetilde{V}_{n})) with a uniform bound of propagation, i.e.,

β„‚HS​[𝐕~]𝝅:={(Tn)∈∏nβˆˆβ„•β„‚HS​[V~n]Ο€n|Β βˆƒR>0Β such thatΒ Prop​(Tn)<RΒ for anyΒ n}\mathbb{C}_{\mathrm{HS}}[\widetilde{\mathbf{V}}]^{\boldsymbol{\pi}}:=\Big{\{}(T_{n})\in\prod_{n\in\mathbb{N}}\mathbb{C}_{\mathrm{HS}}[\widetilde{V}_{n}]^{\pi_{n}}\ \Big{|}\ \text{ $\exists R>0$ such that $\mathrm{Prop}(T_{n})<R$ for any $n$}\Big{\}}

and let Cβˆ—β€‹(𝐕~)𝝅C^{*}(\widetilde{\mathbf{V}})^{\boldsymbol{\pi}} denote its closure in the direct product C*-algebra ∏nβˆˆβ„•Cβˆ—β€‹(V~n)Ο€n\prod_{n\in\mathbb{N}}C^{*}(\widetilde{V}_{n})^{\pi_{n}}. We also define the Real C*-ideal Cβˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅◁Cβˆ—β€‹(𝐕~)𝝅C^{*}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}\triangleleft C^{*}(\widetilde{\mathbf{V}})^{\boldsymbol{\pi}} as the closure of

β„‚HS​[βˆ‚π•~βŠ‚π•~]𝝅:={(Tn)∈∏nβˆˆβ„•β„‚HS​[V~n]Ο€n|Β  >βˆƒR0 such that <⁒Prop(Tn)R and <⁒d(suppTn,βˆ‚Γ—~Vnβˆ‚~Vn)R for any n Β },\mathbb{C}_{\mathrm{HS}}[\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}}]^{\boldsymbol{\pi}}:=\bigg{\{}(T_{n})\in\prod_{n\in\mathbb{N}}\mathbb{C}_{\mathrm{HS}}[\widetilde{V}_{n}]^{\pi_{n}}\ \bigg{|}\ \text{ \parbox{154.99951pt}{$\exists R>0$ such that $\mathrm{Prop}(T_{n})<R$ and $d(\operatorname{\mathrm{supp}}T_{n},\partial\widetilde{V}_{n}\times\partial\widetilde{V}_{n})<R$ for any $n$} }\bigg{\}}_{\textstyle,}

where d​(supp⁑Tn,βˆ‚V~nΓ—βˆ‚V~n)d(\operatorname{\mathrm{supp}}T_{n},\partial\widetilde{V}_{n}\times\partial\widetilde{V}_{n}) stands for the Hausdorff distance.

Let Dβˆ—β€‹(M~)Ξ“D^{*}(\widetilde{M})^{\Gamma} denote the closure of the set of bounded operators on L2​(M~)L^{2}(\widetilde{M}) which is of finite propagation and is pseudo-local, i.e., [T,f]βˆˆπ•‚[T,f]\in\mathbb{K} for any f∈Cc​(M~)f\in C_{c}(\widetilde{M}), with respect to the norm

β€–Tβ€–Dβˆ—β€‹(M~)Ξ“:=supSβˆˆβ„‚HS​[M~]Ξ“βˆ–{0}β€–T​Sβ€–Cβˆ—β€‹(M~)Ξ“β€–Sβ€–Cβˆ—β€‹(M~)Ξ“.\|T\|_{D^{*}(\widetilde{M})^{\Gamma}}:=\sup_{S\in\mathbb{C}_{\mathrm{HS}}[\widetilde{M}]^{\Gamma}\setminus\{0\}}\frac{\|TS\|_{C^{*}(\widetilde{M})^{\Gamma}}}{\|S\|_{C^{*}(\widetilde{M})^{\Gamma}}}.

This is a Real C*-algebra [oyono-oyonoTheoryMaximalRoe2009]*Lemma 2.16, including Cβˆ—β€‹(M~)Ξ“C^{*}(\widetilde{M})^{\Gamma} as a Real C*-ideal. We write the quotient as Qβˆ—β€‹(M~)Ξ“:=Dβˆ—β€‹(M~)Ξ“/Cβˆ—β€‹(M~)Ξ“Q^{*}(\widetilde{M})^{\Gamma}:=D^{*}(\widetilde{M})^{\Gamma}/C^{*}(\widetilde{M})^{\Gamma}. A standard fact in coarse index theory is that the K-group of Qβˆ—β€‹(M~)Ξ“Q^{*}(\widetilde{M})^{\Gamma} is isomorphic to the equivariant K-homology KOβˆ—Ξ“β‘(M~)\operatorname{\mathrm{KO}}^{\Gamma}_{*}(\widetilde{M}). In the same way, we also define the Real C*-algebras Dβˆ—β€‹(𝐕~)𝝅D^{*}(\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}, Qβˆ—β€‹(𝐕~)𝝅Q^{*}(\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}, Dβˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅D^{*}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}, and Qβˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅Q^{*}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}. We only note that Dβˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅D^{*}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}} is the closure of the set of pseudo-local, finite propagation operators which is supported near βˆ‚V\partial V and T​f,f​Tβˆˆπ•‚Tf,fT\in\mathbb{K} for any f∈Cc​(π•βˆ–βˆ‚π•)f\in C_{c}(\mathbf{V}\setminus\partial\mathbf{V}).

2.3.

There is a βˆ—\ast-homomorphism

s:Cβˆ—β€‹(M~)Ξ“β†’Cβˆ—β€‹(𝐕~)𝝅/Cβˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅s\colon C^{*}(\widetilde{M})^{\Gamma}\to C^{*}(\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}/C^{*}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}

constructed in the following way. For Tβˆˆβ„‚HS​[M~]Ξ“T\in\mathbb{C}_{\mathrm{HS}}[\widetilde{M}]^{\Gamma} with Prop​(T)≀CR\mathrm{Prop}(T)\leq C_{R}, we define its lift T~n,Rβˆˆβ„‚HS​[V~n]Ο€n\widetilde{T}_{n,R}\in\mathbb{C}_{\mathrm{HS}}[\widetilde{V}_{n}]^{\pi_{n}} in terms of its kernel function as

(2.4) T~n,R​(x~,y~):={T​(f~n​(x~),f~n​(y~))Β ifΒ d​(x~,y~)<CRΒ andΒ x~,y~∈V~n,R,Β 0Β otherwise.\displaystyle\begin{split}\widetilde{T}_{n,R}(\tilde{x},\tilde{y}):=\begin{cases}T(\tilde{f}_{n}(\tilde{x}),\tilde{f}_{n}(\tilde{y}))&\text{ if $d(\tilde{x},\tilde{y})<C_{R}$ and $\tilde{x},\tilde{y}\in\widetilde{V}_{n,R}$, }\\ 0&\text{ otherwise. }\end{cases}\end{split}

A consequence of LemmaΒ 2.1 is that, for each x~∈V~n,R\tilde{x}\in\widetilde{V}_{n,R} and y∈M~y\in\widetilde{M} with d​(f~n​(x~),y)<CRd(\tilde{f}_{n}(\tilde{x}),y)<C_{R}, there is a unique y~∈f~nβˆ’1​(y)\tilde{y}\in\tilde{f}_{n}^{-1}(y) such that T~n,R​(x~,y~)β‰ 0\widetilde{T}_{n,R}(\tilde{x},\tilde{y})\neq 0. In particular, we have Prop​(T~n,R)=Prop​(T)\mathrm{Prop}(\widetilde{T}_{n,R})=\mathrm{Prop}(T). Moreover, the Ξ“n\Gamma_{n}-invariance of TT implies the Ο€n\pi_{n}-invariance of T~n,R\widetilde{T}_{n,R}.

The assignment T↦T~n,RT\mapsto\widetilde{T}_{n,R} is linear and βˆ—\ast-preserving. Moreover, for T,Sβˆˆβ„‚HS​[M~]Ξ“T,S\in\mathbb{C}_{\mathrm{HS}}[\widetilde{M}]^{\Gamma} with Prop​(T)+Prop​(S)≀CR\mathrm{Prop}(T)+\mathrm{Prop}(S)\leq C_{R}, the lifts T~n,R\widetilde{T}_{n,R}, S~n,R\widetilde{S}_{n,R} and T​S~n,R\widetilde{TS}_{n,R} satisfies that

∫z~∈V~n,RT~n,R​(x~,z~)​T~n,R​(z~,y~)​𝑑volg​(z)=(T​S~)n,R​(x~,y~)\int_{\tilde{z}\in\widetilde{V}_{n,R}}\widetilde{T}_{n,R}(\tilde{x},\tilde{z})\widetilde{T}_{n,R}(\tilde{z},\tilde{y})d\mathrm{vol}_{g}(z)=(\widetilde{TS})_{n,R}(\tilde{x},\tilde{y})

if BCR​(x~)∩BCR​(y~)βŠ‚V~n,RB_{C_{R}}(\tilde{x})\cap B_{C_{R}}(\tilde{y})\subset\widetilde{V}_{n,R}. This implies that T~n,R​S~n,Rβˆ’T​S~n,R\widetilde{T}_{n,R}\widetilde{S}_{n,R}-\widetilde{TS}_{n,R} is supported on the (R+CR)(R+C_{R})-neighborhood of βˆ‚V~nΓ—βˆ‚V~n\partial\widetilde{V}_{n}\times\partial\widetilde{V}_{n}, i.e., the lifting is multiplicative modulo boundary.

Recall that the assignment R↦CRR\mapsto C_{R} is increasing and CRβ†’βˆžC_{R}\to\infty as Rβ†’βˆžR\to\infty. Also, for Rβ€²>R>0R^{\prime}>R>0 and Tβˆˆβ„‚HS​[M~]Ξ“T\in\mathbb{C}_{\mathrm{HS}}[\widetilde{M}]^{\Gamma} with Prop​(T)≀CR\mathrm{Prop}(T)\leq C_{R}, the difference T~n,Rβ€²βˆ’T~n,R\widetilde{T}_{n,R^{\prime}}-\widetilde{T}_{n,R} is supported on the Rβ€²R^{\prime}-neighborhood of βˆ‚V~nΓ—βˆ‚V~n\partial\widetilde{V}_{n}\times\partial\widetilde{V}_{n}. This shows that

s​(T):=(T~n,R)nβˆˆβ„•βˆˆCβˆ—β€‹(𝐕~)𝝅/Cβˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅s(T):=(\widetilde{T}_{n,R})_{n\in\mathbb{N}}\in C^{*}(\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}/C^{*}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}

is well-defined independent of the choice of R>0R>0 with Prop​(T)<CR\mathrm{Prop}(T)<C_{R}. By the above argument, this ss forms a βˆ—\ast-homomorphism from β„‚HS​[M~]Ξ“\mathbb{C}_{\mathrm{HS}}[\widetilde{M}]^{\Gamma}. This extends to the βˆ—\ast-homomorphism from Cβˆ—β€‹(M~)Ξ“C^{*}(\widetilde{M})^{\Gamma} by the maximality of the norm on Cβˆ—β€‹(M~)Ξ“C^{*}(\widetilde{M})^{\Gamma}.

2.5.

The βˆ—\ast-homomorphism ss defined in 2.3 extends to

s:Dβˆ—β€‹(M~)Ξ“β†’Dβˆ—β€‹(𝐕~)𝝅/Dβˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅,s\colon D^{*}(\widetilde{M})^{\Gamma}\to D^{*}(\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}/D^{*}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}},

and hence induces

s:Qβˆ—β€‹(M~)Ξ“β†’Qβˆ—β€‹(𝐕~)𝝅/Qβˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅.s\colon Q^{*}(\widetilde{M})^{\Gamma}\to Q^{*}(\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}/Q^{*}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}.

Indeed, without loss of generality, we may assume that there are Riemannian bands Vnβ€²V_{n}^{\prime} such that VnβŠ‚Vnβ€²V_{n}\subset V_{n}^{\prime}, fnf_{n} extends to an isometric immersion of Vnβ€²V_{n}^{\prime}, and dist​(βˆ‚Vn,βˆ‚Vnβ€²)β‰₯1\mathrm{dist}(\partial V_{n},\partial V_{n}^{\prime})\geq 1. Let us decompose an operator T∈Dβˆ—β€‹(M~)Ξ“T\in D^{*}(\widetilde{M})^{\Gamma} into T=T0+T1T=T^{0}+T^{1}, where Prop​(T0)<C1\mathrm{Prop}(T^{0})<C_{1} and T1∈Cβˆ—β€‹(M~)Ξ“T^{1}\in C^{*}(\widetilde{M})^{\Gamma}. Set

s​(T):=(Ξ n​(T0~)n,1′​Πn)nβˆˆβ„•+s​(T1)∈Dβˆ—β€‹(𝐕~)𝝅/Dβˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅,s(T):=(\Pi_{n}(\widetilde{T^{0}})_{n,1}^{\prime}\Pi_{n})_{n\in\mathbb{N}}+s(T^{1})\in D^{*}(\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}/D^{*}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}},

where (β‹…)n,1β€²(\cdot)_{n,1}^{\prime} denotes the lift (2.4) with respect to Vnβ€²V_{n}^{\prime} and R=1R=1, and Ξ n\Pi_{n} denotes the projection onto L2​(Vn)L^{2}(V_{n}). This ss is well-defined independent of the choice of a decomposition T=T0+T1T=T_{0}+T_{1}, and forms a βˆ—\ast-homomorphism. We omit the detail of the proof, because it is proved completely in the same way as [kubotaCodimensionTransferHigher2021]*Proposition 4.3.

2.3. K-theory and the coarse index of Dirac operators

In the last step, we relate the equivariant coarse index of M~\widetilde{M} with that of βˆ‚V~n\partial\widetilde{V}_{n} through the lifting homomorphism ss constructed above.

We define the ideal C0βˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅C^{*}_{0}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}} of Cβˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅C^{*}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}} consisting of operators (Tn)n(T_{n})_{n} such that β€–Tnβ€–β†’0\|T_{n}\|\to 0 as nβ†’βˆžn\to\infty, and set

Cβ™­βˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅:=Cβˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅/C0βˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅.C^{*}_{\flat}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}:=C^{*}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}/C^{*}_{0}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}.

We also define the Real C*-algebras Dβ™­βˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅D^{*}_{\flat}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}} and Qβ™­βˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅Q^{*}_{\flat}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}} in the same way.

Lemma 2.6.

There are isomorphisms

Ο•:KOβˆ—β‘(Cβ™­βˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅)\displaystyle\phi\colon\operatorname{\mathrm{KO}}_{*}(C^{*}_{\flat}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}) β‰…βˆKOβˆ—β‘(Cβˆ—β€‹(βˆ‚+V~n)Ο€n)⨁KOβˆ—β‘(Cβˆ—β€‹(βˆ‚βˆ’V~n)Ο€n)βŠ•βˆKOβˆ—β‘(Cβˆ—β€‹(βˆ‚βˆ’V~n)Ο€n)⨁KOβˆ—β‘(Cβˆ—β€‹(βˆ‚βˆ’V~n)Ο€n),\displaystyle\cong\frac{\prod\operatorname{\mathrm{KO}}_{*}(C^{*}(\partial_{+}\widetilde{V}_{n})^{\pi_{n}})}{\bigoplus\operatorname{\mathrm{KO}}_{*}(C^{*}(\partial_{-}\widetilde{V}_{n})^{\pi_{n}})}\oplus\frac{\prod\operatorname{\mathrm{KO}}_{*}(C^{*}(\partial_{-}\widetilde{V}_{n})^{\pi_{n}})}{\bigoplus\operatorname{\mathrm{KO}}_{*}(C^{*}(\partial_{-}\widetilde{V}_{n})^{\pi_{n}})},
Ο•:KOβˆ—β‘(Dβ™­βˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅)\displaystyle\phi\colon\operatorname{\mathrm{KO}}_{*}(D^{*}_{\flat}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}) β‰…βˆKOβˆ—β‘(Dβˆ—β€‹(βˆ‚+V~n)Ο€n)⨁KOβˆ—β‘(Dβˆ—β€‹(βˆ‚+V~n)Ο€n)βŠ•βˆKOβˆ—β‘(Dβˆ—β€‹(βˆ‚βˆ’V~n)Ο€n)⨁KOβˆ—β‘(Dβˆ—β€‹(βˆ‚βˆ’V~n)Ο€n),\displaystyle\cong\frac{\prod\operatorname{\mathrm{KO}}_{*}(D^{*}(\partial_{+}\widetilde{V}_{n})^{\pi_{n}})}{\bigoplus\operatorname{\mathrm{KO}}_{*}(D^{*}(\partial_{+}\widetilde{V}_{n})^{\pi_{n}})}\oplus\frac{\prod\operatorname{\mathrm{KO}}_{*}(D^{*}(\partial_{-}\widetilde{V}_{n})^{\pi_{n}})}{\bigoplus\operatorname{\mathrm{KO}}_{*}(D^{*}(\partial_{-}\widetilde{V}_{n})^{\pi_{n}})},
Ο•:KOβˆ—β‘(Qβ™­βˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅)\displaystyle\phi\colon\operatorname{\mathrm{KO}}_{*}(Q^{*}_{\flat}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}) β‰…βˆKOβˆ—β‘(Qβˆ—β€‹(βˆ‚+V~n)Ο€n)⨁KOβˆ—β‘(Qβˆ—β€‹(βˆ‚+V~n)Ο€n)βŠ•βˆKOβˆ—β‘(Qβˆ—β€‹(βˆ‚βˆ’V~n)Ο€n)⨁KOβˆ—β‘(Qβˆ—β€‹(βˆ‚βˆ’V~n)Ο€n).\displaystyle\cong\frac{\prod\operatorname{\mathrm{KO}}_{*}(Q^{*}(\partial_{+}\widetilde{V}_{n})^{\pi_{n}})}{\bigoplus\operatorname{\mathrm{KO}}_{*}(Q^{*}(\partial_{+}\widetilde{V}_{n})^{\pi_{n}})}\oplus\frac{\prod\operatorname{\mathrm{KO}}_{*}(Q^{*}(\partial_{-}\widetilde{V}_{n})^{\pi_{n}})}{\bigoplus\operatorname{\mathrm{KO}}_{*}(Q^{*}(\partial_{-}\widetilde{V}_{n})^{\pi_{n}})}.

We write ϕ±\phi_{\pm} for the first and the second component of Ο•\phi respectively.

Proof.

Let Vn,Β±βŠ‚VnV_{n,\pm}\subset V_{n} denote the n/4n/4-neighborhood of βˆ‚Β±V\partial_{\pm}V and let 𝐕±:=⨆nβˆˆβ„•Vn,Β±\mathbf{V}_{\pm}:=\bigsqcup_{n\in\mathbb{N}}V_{n,\pm}. Then the inclusion

Cβˆ—β€‹(βˆ‚+V~βŠ‚V~+)π…βŠ•Cβˆ—β€‹(βˆ‚βˆ’V~βŠ‚V~βˆ’)𝝅→Cβˆ—β€‹(βˆ‚V~βŠ‚V~)𝝅C^{*}(\partial_{+}\widetilde{\textbf{V}}\subset\widetilde{\textbf{V}}_{+})^{\boldsymbol{\pi}}\oplus C^{*}(\partial_{-}\widetilde{\textbf{V}}\subset\widetilde{\textbf{V}}_{-})^{\boldsymbol{\pi}}\to C^{*}(\partial\widetilde{\textbf{V}}\subset\widetilde{\textbf{V}})^{\boldsymbol{\pi}}

induces a βˆ—\ast-isomorphism

Cβ™­βˆ—β€‹(βˆ‚+V~βŠ‚V~+)π…βŠ•Cβ™­βˆ—β€‹(βˆ‚βˆ’V~βŠ‚V~βˆ’)𝝅→Cβ™­βˆ—β€‹(βˆ‚V~βŠ‚V~)𝝅.C^{*}_{\flat}(\partial_{+}\widetilde{\textbf{V}}\subset\widetilde{\textbf{V}}_{+})^{\boldsymbol{\pi}}\oplus C^{*}_{\flat}(\partial_{-}\widetilde{\textbf{V}}\subset\widetilde{\textbf{V}}_{-})^{\boldsymbol{\pi}}\to C^{*}_{\flat}(\partial\widetilde{\textbf{V}}\subset\widetilde{\textbf{V}})^{\boldsymbol{\pi}}.

Now the six-term exact sequence for the extension

0β†’C0βˆ—β€‹(βˆ‚Β±π•~βŠ‚π•~Β±)𝝅→Cβˆ—β€‹(βˆ‚Β±π•~βŠ‚π•~Β±)𝝅→Cβ™­βˆ—β€‹(βˆ‚Β±π•~βŠ‚π•~Β±)𝝅→00\to C^{*}_{0}(\partial_{\pm}\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}}_{\pm})^{\boldsymbol{\pi}}\to C^{*}(\partial_{\pm}\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}}_{\pm})^{\boldsymbol{\pi}}\to C^{*}_{\flat}(\partial_{\pm}\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}}_{\pm})^{\boldsymbol{\pi}}\to 0

proves the first isomorphism. The second and the third isomorphisms are also proved in the same way. ∎

Let

(2.7) βˆ‚:KOβˆ—β‘(Cβˆ—β€‹(𝐕~)𝝅Cβˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅)β†’KOβˆ—βˆ’1⁑(Cβ™­βˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅)\displaystyle\partial\colon\operatorname{\mathrm{KO}}_{*}\Big{(}\frac{C^{*}(\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}}{C^{*}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}}\Big{)}\to\operatorname{\mathrm{KO}}_{*-1}(C^{*}_{\flat}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}})

denote the K-theory boundary map associated to the exact sequence

0β†’Cβˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅C0βˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅→Cβˆ—β€‹(𝐕~)𝝅C0βˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅→Cβˆ—β€‹(𝐕~)𝝅Cβˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅→0.0\to\frac{C^{*}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}}{C^{*}_{0}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}}\to\frac{C^{*}(\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}}{C^{*}_{0}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}}\to\frac{C^{*}(\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}}{C^{*}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}}\to 0.

We also use the same letter βˆ‚\partial for the KO\operatorname{\mathrm{KO}}-theory boundary map of the same kind defined for Dβˆ—D^{*} and Qβˆ—Q^{*} coarse C*-algebras.

Let DΜΈM~\not{D}_{\widetilde{M}}, DΜΈV~n\not{D}_{\widetilde{V}_{n}}, DΜΈβˆ‚Β±V~n\not{D}_{\partial_{\pm}\widetilde{V}_{n}} denote the Cβ„“d,0\text{\it C$\ell$}_{d,0}-linear Dirac operator on M~\widetilde{M} and V~n\widetilde{V}_{n}, and the Cβ„“dβˆ’1,0\text{\it C$\ell$}_{d-1,0}-linear Dirac operator on βˆ‚Β±V~n\partial_{\pm}\widetilde{V}_{n} respectively (for the definition of Cβ„“d,0\text{\it C$\ell$}_{d,0}-linear Dirac operator on a dd-dimensional spin manifold, see e.g.Β [lawsonjr.SpinGeometry1989]*Chapter II, Β§7). We also consider the Clifford-linear Dirac operators D̸𝐕~\not{D}_{\widetilde{\mathbf{V}}} and DΜΈβˆ‚Β±π•~\not{D}_{\partial_{\pm}\widetilde{\mathbf{V}}}, each of which is the same thing as the family (DΜΈV~n)nβˆˆβ„•(\not{D}_{\widetilde{V}_{n}})_{n\in\mathbb{N}} and (DΜΈβˆ‚Β±V~n)nβˆˆβ„•(\not{D}_{\partial_{\pm}\widetilde{V}_{n}})_{n\in\mathbb{N}} respectively. Let Ο‡:ℝ→ℝ\chi\colon\mathbb{R}\to\mathbb{R} is a continuous function such that χ​(t)β†’Β±1\chi(t)\to\pm 1 as tβ†’Β±βˆžt\to\pm\infty and the support of Ο‡^\hat{\chi} is in (βˆ’1,1)(-1,1). Then χ​(DΜΈM~)\chi(\not{D}_{\widetilde{M}}) is an odd self-adjoint operator in Dβˆ—β€‹(M~)Ξ“D^{*}(\widetilde{M})^{\Gamma} by [roePartitioningNoncompactManifolds1988]*Proposition 2.3, whose image in Qβˆ—β€‹(M~)Ξ“Q^{*}(\widetilde{M})^{\Gamma} is unitary. Hence it determines an element of KOd+1⁑(Qβˆ—β€‹(M~)Ξ“)\operatorname{\mathrm{KO}}_{d+1}(Q^{*}(\widetilde{M})^{\Gamma}) (cf.Β [kubotaCodimensionTransferHigher2021]*Remark A.1), which is denoted by [DΜΈM~][\not{D}_{\widetilde{M}}] in short. Similarly, we define [DΜΈβˆ‚Β±V~n]∈KOd⁑(Qβˆ—β€‹(βˆ‚Β±V~n)Ο€n)[\not{D}_{\partial_{\pm}\widetilde{V}_{n}}]\in\operatorname{\mathrm{KO}}_{d}(Q^{*}(\partial_{\pm}\widetilde{V}_{n})^{\pi_{n}}), and [DΜΈβˆ‚π•~]∈KOd⁑(Qβˆ—β€‹(βˆ‚π•~)𝝅)[\not{D}_{\partial\widetilde{\mathbf{V}}}]\in\operatorname{\mathrm{KO}}_{d}(Q^{*}(\partial\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}). We also define the relative KO-class of the Dirac operators on manifolds with boundary [DΜΈV~n]∈KOd+1⁑(Qβˆ—β€‹(V~n)Ο€n/Qβˆ—β€‹(βˆ‚V~nβŠ‚V~n)Ο€n)[\not{D}_{\widetilde{V}_{n}}]\in\operatorname{\mathrm{KO}}_{d+1}(Q^{*}(\widetilde{V}_{n})^{\pi_{n}}/Q^{*}(\partial\widetilde{V}_{n}\subset\widetilde{V}_{n})^{\pi_{n}}) and [D̸𝐕~]∈KOd+1⁑(Qβˆ—β€‹(𝐕~)𝝅/Qβˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅)[\not{D}_{\widetilde{\mathbf{V}}}]\in\operatorname{\mathrm{KO}}_{d+1}(Q^{*}(\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}/Q^{*}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}).

The K-theory boundary map of the extensions 0β†’Cβˆ—β€‹(X)Gβ†’Dβˆ—β€‹(X)Gβ†’Qβˆ—β€‹(X)Gβ†’00\to C^{*}(X)^{G}\to D^{*}(X)^{G}\to Q^{*}(X)^{G}\to 0 are denoted by IndG\operatorname{\mathrm{Ind}}_{G}, where (X,G)(X,G) is (M~,Ξ“)(\widetilde{M},\Gamma), (Vn,Ο€n)(V_{n},\pi_{n}), (𝐕~,𝝅)(\widetilde{\mathbf{V}},\boldsymbol{\pi}), and so on. Note that Cβˆ—β€‹(M)Ξ“C^{*}(M)^{\Gamma} is Morita equivalent to the maximal group C*-algebra Cβˆ—β€‹Ξ“C^{*}\Gamma and the Rosenberg index αΓ​(M)\alpha_{\Gamma}(M) is the same thing with IndΓ⁑([DΜΈM~])\operatorname{\mathrm{Ind}}_{\Gamma}([\not{D}_{\widetilde{M}}]).

Hereafter, for a sequence of abelian groups AnA_{n} and an element (an)n∈∏An(a_{n})_{n}\in\prod A_{n}, we write (an)nβ™­(a_{n})_{n}^{\flat} for its image by the quotient map ∏Anβ†’βˆAn/⨁An\prod A_{n}\to\prod A_{n}/\bigoplus A_{n}.

Lemma 2.8.

The composition

KOd+1⁑(Qβˆ—β€‹(M~)Ξ“)β†’sβˆ—KOd+1⁑(Qβˆ—β€‹(𝐕~)𝝅Qβˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅)β†’βˆ‚KOd⁑(Qβ™­βˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅)β†’Ο•+∏KOd⁑(Qβˆ—β€‹(βˆ‚+V~n)Ο€n)⨁KOd⁑(Qβˆ—β€‹(βˆ‚+V~n)Ο€n)\operatorname{\mathrm{KO}}_{d+1}(Q^{*}(\widetilde{M})^{\Gamma})\xrightarrow{s_{*}}\operatorname{\mathrm{KO}}_{d+1}\Big{(}\frac{Q^{*}(\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}}{Q^{*}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}}\Big{)}\xrightarrow{\partial}\operatorname{\mathrm{KO}}_{d}(Q^{*}_{\flat}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}})\xrightarrow{\phi_{+}}\frac{\prod\operatorname{\mathrm{KO}}_{d}(Q^{*}(\partial_{+}\widetilde{V}_{n})^{\pi_{n}})}{\bigoplus\operatorname{\mathrm{KO}}_{d}(Q^{*}(\partial_{+}\widetilde{V}_{n})^{\pi_{n}})}

sends [DΜΈM~][\not{D}_{\widetilde{M}}] to ([DΜΈβˆ‚+V~n])nβ™­([\not{D}_{\partial_{+}\widetilde{V}_{n}}])_{n}^{\flat}.

Proof.

Firstly, we have sβˆ—β€‹[DΜΈM~]=[D̸𝐕~]s_{*}[\not{D}_{\widetilde{M}}]=[\not{D}_{\widetilde{\mathbf{V}}}]. This is because the operators s​(χ​(DΜΈM~))s(\chi(\not{D}_{\widetilde{M}})) and χ​(D̸𝐕)\chi(\not{D}_{\mathbf{V}}) are both 0-th order pseudo-differential operators and their principal symbols are the same. Next, βˆ‚[D̸𝐕~]=[DΜΈβˆ‚+𝐕~]\partial[\not{D}_{\widetilde{\mathbf{V}}}]=[\not{D}_{\partial_{+}\widetilde{\mathbf{V}}}] follows from the β€˜boundary of Dirac is Dirac’ principle (see e.g.Β [higsonAnalyticHomology2000]*Proposition 11.2.15). Finally, Ο•+​([DΜΈβˆ‚+𝐕~])=([DΜΈβˆ‚+V~n])nβ™­\phi_{+}([\not{D}_{\partial_{+}\widetilde{\mathbf{V}}}])=([\not{D}_{\partial_{+}\widetilde{V}_{n}}])_{n}^{\flat} is obvious from the definition. ∎

Proof of TheoremΒ 1.3.

By definition, the diagram

KOd+1⁑(Qβˆ—β€‹(M~)Ξ“)\textstyle{\operatorname{\mathrm{KO}}_{d+1}(Q^{*}(\widetilde{M})^{\Gamma})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}sβˆ—\scriptstyle{s_{*}}IndΞ“\scriptstyle{\operatorname{\mathrm{Ind}}_{\Gamma}}KOd+1⁑(Qβˆ—β€‹(𝐕~)𝝅Qβˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅)\textstyle{\operatorname{\mathrm{KO}}_{d+1}\Big{(}\frac{Q^{*}(\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}}{Q^{*}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}}\Big{)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}βˆ‚\scriptstyle{\partial}Ind𝝅\scriptstyle{\operatorname{\mathrm{Ind}}_{\boldsymbol{\pi}}}KOd⁑(Qβ™­βˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅)\textstyle{\operatorname{\mathrm{KO}}_{d}(Q^{*}_{\flat}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ind𝝅\scriptstyle{\operatorname{\mathrm{Ind}}_{\boldsymbol{\pi}}}Ο•+\scriptstyle{\phi_{+}}∏KOd⁑(Qβˆ—β€‹(βˆ‚+V~n)Ο€n)⨁KOd⁑(Qβˆ—β€‹(βˆ‚+V~n)Ο€n)\textstyle{\frac{\prod\operatorname{\mathrm{KO}}_{d}(Q^{*}(\partial_{+}\widetilde{V}_{n})^{\pi_{n}})}{\bigoplus\operatorname{\mathrm{KO}}_{d}(Q^{*}(\partial_{+}\widetilde{V}_{n})^{\pi_{n}})}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(IndΟ€n)nβ™­\scriptstyle{(\operatorname{\mathrm{Ind}}_{\pi_{n}})_{n}^{\flat}}KOd⁑(Cβˆ—β€‹(M~)Ξ“)\textstyle{\operatorname{\mathrm{KO}}_{d}(C^{*}(\widetilde{M})^{\Gamma})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}sβˆ—\scriptstyle{s_{*}}KOd⁑(Cβˆ—β€‹(𝐕~)𝝅Cβˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅)\textstyle{\operatorname{\mathrm{KO}}_{d}\Big{(}\frac{C^{*}(\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}}{C^{*}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}}\Big{)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}βˆ‚\scriptstyle{\partial}KOdβˆ’1⁑(Cβ™­βˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅)\textstyle{\operatorname{\mathrm{KO}}_{d-1}(C^{*}_{\flat}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ο•+\scriptstyle{\phi_{+}}∏KOdβˆ’1⁑(Cβˆ—β€‹(βˆ‚+V~n)Ο€n)⨁KOdβˆ’1⁑(Cβˆ—β€‹(βˆ‚+V~n)Ο€n)\textstyle{\frac{\prod\operatorname{\mathrm{KO}}_{d-1}(C^{*}(\partial_{+}\widetilde{V}_{n})^{\pi_{n}})}{\bigoplus\operatorname{\mathrm{KO}}_{d-1}(C^{*}(\partial_{+}\widetilde{V}_{n})^{\pi_{n}})}}

commutes. Therefore, we get

(Ο•+βˆ˜βˆ‚βˆ˜sβˆ—)​(IndΓ⁑([DΜΈM~]))\displaystyle(\phi_{+}\circ\partial\circ s_{*})(\operatorname{\mathrm{Ind}}_{\Gamma}([\not{D}_{\widetilde{M}}])) =(IndΟ€n)n♭​((Ο•+βˆ˜βˆ‚βˆ˜sβˆ—)​([DΜΈM~]))=(IndΟ€n⁑([DΜΈβˆ‚+V~n]))nβ™­.\displaystyle=(\operatorname{\mathrm{Ind}}_{\pi_{n}})_{n}^{\flat}((\phi_{+}\circ\partial\circ s_{*})([\not{D}_{\widetilde{M}}]))=(\operatorname{\mathrm{Ind}}_{\pi_{n}}([\not{D}_{\partial_{+}\widetilde{V}_{n}}]))_{n}^{\flat}.

The right hand side is non-zero by assumption. This shows the non-vanishing of IndΓ⁑([DΜΈM~])\operatorname{\mathrm{Ind}}_{\Gamma}([\not{D}_{\widetilde{M}}]) as desired. ∎

3. Multiwidth of cube-like domains and the Rosenberg index

In this section we study a β€˜multi-dimensional’ generalization of TheoremΒ 1.3.

Definition 3.1.

A dd-dimensional Riemannian β–‘m\square^{m}-domain is a compact Riemannian manifold (V,g)(V,g) with corners equipped with a face-preserving smooth map f:Vβ†’[βˆ’1,1]mf\colon V\to[-1,1]^{m} which is corner proper, i.e., 11-faces of VV are pull-backs of 11-faces of [βˆ’1,1]m[-1,1]^{m} (cf.Β [gromovFourLecturesScalar2019]*3.18).

In this terminology, β–‘1\square^{1}-domain is the same thing as Riemannian band.

We write the codimension 11 faces of VV as βˆ‚j,Β±V:=fβˆ’1​(pjβˆ’1​({Β±1}))\partial_{j,\pm}V:=f^{-1}(p_{j}^{-1}(\{\pm 1\})). Note that the codimension mm corner

Vβ‹”:=βˆ‚1,+Vβˆ©β‹―βˆ©βˆ‚m,+VV_{\pitchfork}:=\partial_{1,+}V\cap\cdots\cap\partial_{m,+}V

is a closed manifold.

Definition 3.2.

We define the multiwidth of a β–‘m\square^{m}-domain as

width​(V,g):=minj=1,…,m⁑dist​(βˆ‚j,+V,βˆ‚j,βˆ’V).\mathrm{width}(V,g):=\min_{j=1,\dots,m}\mathrm{dist}(\partial_{j,+}V,\partial_{j,-}V).

For a class 𝒱m\mathcal{V}_{m} of Riemannian β–‘m\square^{m}-domains, we define the 𝒱m\mathcal{V}_{m}-multiwidth of a complete spin manifold MM, denoted by width𝒱m​(M,g)\mathrm{width}_{\mathcal{V}_{m}}(M,g), as the supremum of the width of a Riemannian β–‘nβˆ’m\square^{n-m}-domains immersed to MM.

Definition 3.3.

We say that a β–‘m\square^{m}-domain VV is in the class 𝒦​π’ͺm\mathcal{KO}_{m} if the equivariant coarse index

IndΟ€1​(V)⁑(DΜΈV~β‹”)∈KOdβˆ’m⁑(Cβˆ—β€‹(V~β‹”)Ο€1​(V))\operatorname{\mathrm{Ind}}_{\pi_{1}(V)}(\not{D}_{\widetilde{V}_{\pitchfork}})\in\operatorname{\mathrm{KO}}_{d-m}(C^{*}(\widetilde{V}_{\pitchfork})^{\pi_{1}(V)})

does not vanish.

Remark 3.4.

We compare our assumption on cube-like domains with the previous works [gromovFourLecturesScalar2019, wangProofGromovCube2021, xieQuantitativeRelativeIndex2021]. First, the above papers deal with a manifold with boundary XX, instead of a manifold with corner, equipped with a map f:Xβ†’[βˆ’1,1]mf\colon X\to[-1,1]^{m} sending βˆ‚X\partial X to the boundary of the cube. For such XX, the inverse image Xβ€²:=fβˆ’1​([βˆ’1+Ξ΅,1βˆ’Ξ΅]m)X^{\prime}:=f^{-1}([-1+\varepsilon,1-\varepsilon]^{m}) is a β–‘m\square^{m}-domain in the sense of DefinitionΒ 3.1 if Ξ΅>0\varepsilon>0 is chosen to be a regular value of pj∘fp_{j}\circ f for j=1,…,mj=1,\dots,m (where pjp_{j} denotes the jj-th projection). We may choose such Ξ΅>0\varepsilon>0 in the way that the distance of βˆ‚X\partial X and βˆ‚Xβ€²\partial X^{\prime} is arbitrarily small.

Next, the submanifold playing the role of βˆ‚+V\partial_{+}V in the band width theory in the previous papers is not the lowest dimensional corner Vβ‹”V_{\pitchfork}, but the transverse intersection Yβ‹”Y_{\pitchfork} of mm hypersurfaces YjβŠ‚VY_{j}\subset V which separates βˆ‚j,+V\partial_{j,+}V and βˆ‚j,βˆ’V\partial_{j,-}V. In particular, the assumption of [wangProofGromovCube2021, xieQuantitativeRelativeIndex2021] is the non-vanishing of the Rosenberg index of Yβ‹”Y_{\pitchfork} under the condition that Ο€1​(Yβ‹”)β†’Ο€1​(V)\pi_{1}(Y_{\pitchfork})\to\pi_{1}(V) is injective. Indeed, this assumption is reduced to the non-vanishing of the Ο€1​(V)\pi_{1}(V)-equivariant coarse index of the Ο€1​(V)\pi_{1}(V)-Galois covering Y~β‹”\widetilde{Y}_{\pitchfork} (instead of Ο€1​(Yβ‹”)\pi_{1}(Y_{\pitchfork})-equivariant coarse index of the universal covering of Yβ‹”Y_{\pitchfork}). This reduction is discussed in [xieQuantitativeRelativeIndex2021], in the second paragraph of the proof of Theorem 4.3. Since Yβ‹”Y_{\pitchfork} and Vβ‹”V_{\pitchfork} are cobordant, their equivariant coarse index coincides.

Let MM be a closed spin manifold. Assume that MM has infinite 𝒦​π’ͺm\mathcal{KO}_{m}-multiwidth. For nβˆˆβ„•n\in\mathbb{N}, pick a Riemannian β–‘m\square^{m}-domain VnV_{n} in the class 𝒦​π’ͺm\mathcal{KO}_{m} which has the multiwidth >n>n and is isometrically immersed to MM. We use the letters Ο€n\pi_{n}, Ξ“n\Gamma_{n}, 𝐕\mathbf{V}, and 𝐕~\widetilde{\mathbf{V}} in the same way as SectionΒ 2. Moreover, we define the maximal Roe algebras Cβˆ—β€‹(𝐕~)𝝅C^{*}(\widetilde{\mathbf{V}})^{\boldsymbol{\pi}} and Cβˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅C^{*}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}} in the same way. Note that, the same proof as LemmaΒ 2.1 shows that, the image of V~n,R\widetilde{V}_{n,R} in M~\widetilde{M} has systole not less than 2​CR2C_{R}. This means that the same construction as 2.3 and 2.5 works, and hence we get the βˆ—\ast-homomorphisms

s\displaystyle s :Cβˆ—β€‹(M~)Ξ“β†’Cβˆ—β€‹(𝐕~)𝝅/Cβˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅,\displaystyle\colon C^{*}(\widetilde{M})^{\Gamma}\to C^{*}(\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}/C^{*}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}},
s\displaystyle s :Dβˆ—β€‹(M~)Ξ“β†’Dβˆ—β€‹(𝐕~)𝝅/Cβˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅,\displaystyle\colon D^{*}(\widetilde{M})^{\Gamma}\to D^{*}(\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}/C^{*}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}},
s\displaystyle s :Qβˆ—β€‹(M~)Ξ“β†’Qβˆ—β€‹(𝐕~)𝝅/Cβˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅.\displaystyle\colon Q^{*}(\widetilde{M})^{\Gamma}\to Q^{*}(\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}/C^{*}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}.

Let Wn:=βˆ‚1,+VW_{n}:=\partial_{1,+}V and let W~n\widetilde{W}_{n} denote the Ο€n\pi_{n}-Galois covering of WnW_{n} (note that it is not necessarily the universal covering of WnW_{n}). Set 𝐖=⨆nβˆˆβ„•Wn\mathbf{W}=\bigsqcup_{n\in\mathbb{N}}W_{n} and 𝐖~=⨆nβˆˆβ„•W~n\widetilde{\mathbf{W}}=\bigsqcup_{n\in\mathbb{N}}\widetilde{W}_{n}. We remark that WnW_{n} is a Riemannian β–‘mβˆ’1\square^{m-1}-domain.

Let D̸𝐕~\not{D}_{\widetilde{\mathbf{V}}} and D̸𝐖~\not{D}_{\widetilde{\mathbf{W}}} denote the Dirac operator on 𝐕~\widetilde{\mathbf{V}} and 𝐖~\widetilde{\mathbf{W}} respectively. In the same way as the previous section, these operators determine the KO\operatorname{\mathrm{KO}}-classes [D̸𝐕~]∈KOd+1⁑(Qβˆ—β€‹(𝐕~)𝝅/Qβˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅)[\not{D}_{\widetilde{\mathbf{V}}}]\in\operatorname{\mathrm{KO}}_{d+1}(Q^{*}(\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}/Q^{*}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}) and [D̸𝐖~]∈KOd⁑(Qβˆ—β€‹(𝐖~)𝝅/Qβˆ—β€‹(βˆ‚π–~βŠ‚π–~)𝝅)[\not{D}_{\widetilde{\mathbf{W}}}]\in\operatorname{\mathrm{KO}}_{d}(Q^{*}(\widetilde{\mathbf{W}})^{\boldsymbol{\pi}}/Q^{*}(\partial\widetilde{\mathbf{W}}\subset\widetilde{\mathbf{W}})^{\boldsymbol{\pi}}) respectively.

Lemma 3.5.

There is a homomorphism

βˆ‚:KOd⁑(Cβˆ—β€‹(𝐕~)𝝅Cβˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅)β†’KOdβˆ’1⁑(Cβˆ—β€‹(𝐖~)𝝅Cβˆ—β€‹(βˆ‚π–~βŠ‚π–~)𝝅)\displaystyle\partial\colon\operatorname{\mathrm{KO}}_{d}\bigg{(}\frac{C^{*}(\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}}{C^{*}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}}\bigg{)}\to\operatorname{\mathrm{KO}}_{d-1}\bigg{(}\frac{C^{*}(\widetilde{\mathbf{W}})^{\boldsymbol{\pi}}}{C^{*}(\partial\widetilde{\mathbf{W}}\subset\widetilde{\mathbf{W}})^{\boldsymbol{\pi}}}\bigg{)}

sending Ind𝛑⁑([D̸𝐕~])\operatorname{\mathrm{Ind}}_{\boldsymbol{\pi}}([\mathbf{\not{D}}_{\widetilde{\mathbf{V}}}]) to Ind𝛑⁑([D̸𝐖~])\operatorname{\mathrm{Ind}}_{\boldsymbol{\pi}}([\mathbf{\not{D}}_{\widetilde{\mathbf{W}}}]).

Proof.

Let 𝐙~\widetilde{\mathbf{Z}} denote the closure of βˆ‚π•~βˆ–π–~\partial\widetilde{\mathbf{V}}\setminus\widetilde{\mathbf{W}}. Note that 𝐙~βˆ©π–~=βˆ‚π–~\widetilde{\mathbf{Z}}\cap\widetilde{\mathbf{W}}=\partial\widetilde{\mathbf{W}}, and hence

Cβˆ—β€‹(βˆ‚π•~)𝝅/Cβˆ—β€‹(𝐙~βŠ‚βˆ‚π•~)𝝅≅Cβˆ—β€‹(𝐖~)𝝅/Cβˆ—β€‹(βˆ‚π–~βŠ‚π–~)𝝅.C^{*}(\partial\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}/C^{*}(\widetilde{\mathbf{Z}}\subset\partial\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}\cong C^{*}(\widetilde{\mathbf{W}})^{\boldsymbol{\pi}}/C^{*}(\partial\widetilde{\mathbf{W}}\subset\widetilde{\mathbf{W}})^{\boldsymbol{\pi}}.

The Real C*-algebra extension

0β†’Cβˆ—β€‹(βˆ‚π•~)𝝅Cβˆ—β€‹(𝐙~βŠ‚βˆ‚π•~)𝝅→Cβˆ—β€‹(𝐕~)𝝅Cβˆ—β€‹(𝐙~βŠ‚π•~)𝝅→Cβˆ—β€‹(𝐕~)𝝅Cβˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅→00\to\frac{C^{*}(\partial\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}}{C^{*}(\widetilde{\mathbf{Z}}\subset\partial\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}}\to\frac{C^{*}(\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}}{C^{*}(\widetilde{\mathbf{Z}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}}\to\frac{C^{*}(\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}}{C^{*}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}}\to 0

induces the boundary map

βˆ‚:KOd⁑(Cβˆ—β€‹(𝐕~)𝝅Cβˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅)β†’KOdβˆ’1⁑(Cβˆ—β€‹(βˆ‚π•~)Cβˆ—β€‹(βˆ‚π™~βŠ‚βˆ‚π•~))β‰…KOdβˆ’1⁑(Cβˆ—β€‹(𝐖~)𝝅Cβˆ—β€‹(βˆ‚π–~βŠ‚π–~)𝝅)\partial\colon\operatorname{\mathrm{KO}}_{d}\Big{(}\frac{C^{*}(\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}}{C^{*}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}}\Big{)}\to\operatorname{\mathrm{KO}}_{d-1}\Big{(}\frac{C^{*}(\partial\widetilde{\mathbf{V}})}{C^{*}(\partial\widetilde{\mathbf{Z}}\subset\partial\widetilde{\mathbf{V}})}\Big{)}\cong\operatorname{\mathrm{KO}}_{d-1}\Big{(}\frac{C^{*}(\widetilde{\mathbf{W}})^{\boldsymbol{\pi}}}{C^{*}(\partial\widetilde{\mathbf{W}}\subset\widetilde{\mathbf{W}})^{\boldsymbol{\pi}}}\Big{)}

as desired. Moreover, the boundary of the same kind are also defined for Dβˆ—D^{*} and Qβˆ—Q^{*} coarse C*-algebras in the same way. Now, the commutativity of the diagram

KOd+1⁑(Qβˆ—β€‹(𝐕~)𝝅Qβˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅)\textstyle{\operatorname{\mathrm{KO}}_{d+1}\Big{(}\frac{Q^{*}(\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}}{Q^{*}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}}\Big{)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}βˆ‚\scriptstyle{\partial}Ind𝝅\scriptstyle{\operatorname{\mathrm{Ind}}_{\boldsymbol{\pi}}}KOd⁑(Qβˆ—β€‹(𝐕~)𝝅Qβˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅)\textstyle{\operatorname{\mathrm{KO}}_{d}\Big{(}\frac{Q^{*}(\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}}{Q^{*}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}}\Big{)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ind𝝅\scriptstyle{\operatorname{\mathrm{Ind}}_{\boldsymbol{\pi}}}KOd⁑(Cβˆ—β€‹(𝐕~)𝝅Cβˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅)\textstyle{\operatorname{\mathrm{KO}}_{d}\Big{(}\frac{C^{*}(\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}}{C^{*}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}}\Big{)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}βˆ‚\scriptstyle{\partial}KOdβˆ’1⁑(Cβˆ—β€‹(𝐕~)𝝅Cβˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅)\textstyle{\operatorname{\mathrm{KO}}_{d-1}\Big{(}\frac{C^{*}(\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}}{C^{*}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}}\Big{)}}

and the boundary of Dirac is Dirac principle, βˆ‚[D̸𝐕]=[D̸𝐖]\partial[\not{D}_{\mathbf{V}}]=[\not{D}_{\mathbf{W}}], shows that βˆ‚(Ind𝝅⁑([D̸𝐕~]))=Ind𝝅⁑([D̸𝐖~])\partial(\operatorname{\mathrm{Ind}}_{\boldsymbol{\pi}}([\mathbf{\not{D}}_{\widetilde{\mathbf{V}}}]))=\operatorname{\mathrm{Ind}}_{\boldsymbol{\pi}}([\mathbf{\not{D}}_{\widetilde{\mathbf{W}}}]) as desired. ∎

Proof of TheoremΒ 1.5.

For 1≀k≀m1\leq k\leq m, let WnkW_{n}^{k} denote the face βˆ‚1,+Vβˆ©β‹―β€‹βˆ‚k,+V\partial_{1,+}V\cap\cdots\partial_{k,+}V, which is a β–‘mβˆ’k\square^{m-k}-domain. Let us consider the mβˆ’1m-1 iterated composition of the map defined in LemmaΒ 3.5 as

KOd(Cβˆ—β€‹(𝐕~)𝝅Cβˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅)β†’βˆ‚KOdβˆ’1(Cβˆ—β€‹(𝐖1~)𝝅Cβˆ—β€‹(βˆ‚π–~1βŠ‚π–~1)𝝅)β†’βˆ‚β‹―β†’βˆ‚KOdβˆ’m+1(Cβˆ—β€‹(𝐖~mβˆ’1)𝝅Cβˆ—β€‹(βˆ‚π–~mβˆ’1βŠ‚π–~mβˆ’1)𝝅).\mathclap{\operatorname{\mathrm{KO}}_{d}\Big{(}\frac{C^{*}(\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}}{C^{*}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}}\Big{)}\xrightarrow{\partial}\operatorname{\mathrm{KO}}_{d-1}\Big{(}\frac{C^{*}(\widetilde{\mathbf{W}^{1}})^{\boldsymbol{\pi}}}{C^{*}(\partial\widetilde{\mathbf{W}}^{1}\subset\widetilde{\mathbf{W}}^{1})^{\boldsymbol{\pi}}}\Big{)}\xrightarrow{\partial}\cdots\xrightarrow{\partial}\operatorname{\mathrm{KO}}_{d-m+1}\Big{(}\frac{C^{*}(\widetilde{\mathbf{W}}^{m-1})^{\boldsymbol{\pi}}}{C^{*}(\partial\widetilde{\mathbf{W}}^{m-1}\subset\widetilde{\mathbf{W}}^{m-1})^{\boldsymbol{\pi}}}\Big{)}_{\textstyle.}}

We further compose the boundary map (2.7) and Ο•+\phi_{+}. Finally we get

Ο•+βˆ˜βˆ‚m:KOd⁑(Cβˆ—β€‹(𝐕~)𝝅Cβˆ—β€‹(βˆ‚π•~βŠ‚π•~)𝝅)β†’βˆKOdβˆ’m⁑(Cβˆ—β€‹(V~n,β‹”)Ο€n)⨁KOdβˆ’m⁑(Cβˆ—β€‹(V~n,β‹”)Ο€n).\phi_{+}\circ\partial^{m}\colon\operatorname{\mathrm{KO}}_{d}\Big{(}\frac{C^{*}(\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}}{C^{*}(\partial\widetilde{\mathbf{V}}\subset\widetilde{\mathbf{V}})^{\boldsymbol{\pi}}}\Big{)}\to\frac{\prod\operatorname{\mathrm{KO}}_{d-m}(C^{*}(\widetilde{V}_{n,\pitchfork})^{\pi_{n}})}{\bigoplus\operatorname{\mathrm{KO}}_{d-m}(C^{*}(\widetilde{V}_{n,\pitchfork})^{\pi_{n}})}_{\textstyle.}

By LemmaΒ 3.5 and the proof of TheoremΒ 1.3, this map sends the equivariant coarse index Ind𝝅⁑(D̸𝐕~)\operatorname{\mathrm{Ind}}_{\boldsymbol{\pi}}(\not{D}_{\widetilde{\mathbf{V}}}) to (IndΟ€n⁑(DΜΈV~n,β‹”))nβ™­(\operatorname{\mathrm{Ind}}_{\pi_{n}}(\not{D}_{\widetilde{V}_{n,\pitchfork}}))_{n}^{\flat}, which does not vanish by assumption. This finishes the proof. ∎

References