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

Higher localised A^\widehat{A}-genera for proper actions and applications

Hao Guo Yau Mathematical Sciences Center, Tsinghua University [email protected]  and  Varghese Mathai School of Mathematical Sciences, University of Adelaide [email protected]
Abstract.

For a finitely generated discrete group Γ\Gamma acting properly on a spin manifold MM, we formulate new topological obstructions to Γ\Gamma-invariant metrics of positive scalar curvature on MM that take into account the cohomology of the classifying space B¯Γ\underline{B}\Gamma for proper actions.

In the cocompact case, this leads to a natural generalisation of Gromov-Lawson’s notion of higher A^\widehat{A}-genera to the setting of proper actions by groups with torsion. It is conjectured that these invariants obstruct the existence of Γ\Gamma-invariant positive scalar curvature on MM. For classes arising from the subring of H(B¯Γ,)H^{*}(\underline{B}\Gamma,\mathbb{R}) generated by elements of degree at most 22, we are able to prove this, under suitable assumptions, using index-theoretic methods for projectively invariant Dirac operators and a twisted L2L^{2}-Lefschetz fixed-point theorem involving a weighted trace on conjugacy classes. The latter generalises a result of Wang-Wang [24] to the projective setting. In the special case of free actions and the trivial conjugacy class, this reduces to a theorem of Mathai [17], which provided a partial answer to a conjecture of Gromov-Lawson on higher A^\widehat{A}-genera.

If MM is non-cocompact, we obtain obstructions to MM being a partitioning hypersurface inside a non-cocompact Γ\Gamma-manifold with non-negative scalar curvature that is positive in a neighbourhood of the hypersurface. Finally, we define a quantitative version of the twisted higher index, as first introduced in [12], and use it to prove a parameterised vanishing theorem in terms of the lower bound of the total curvature term in the square of the twisted Dirac operator.

Key words and phrases:
Higher index, positive scalar curvature, fixed point theorem, coarse geometry, twisted Roe algebra, quantitative index
2010 Mathematics Subject Classification:
46L80, 58B34, 53C20
H.G. and V.M. were partially supported by funding from the Australian Research Council, through the Discovery Project Grant DP200100729. V.M. was supported by the Australian Research Council, through the Australian Laureate Fellowship FL170100020. H.G. was partially supported by NSF DMS-2000082. The authors thank Hang Wang, Zhizhang Xie, and Guoliang Yu for their helpful comments. H.G. is grateful for useful feedback from the audience at the 2021 NUS Conference on Index Theory and Related Topics, August 21-25, 2021

1. Introduction

In this paper we develop the theory of higher indices of projectively equivariant Dirac operators with respect to proper actions of discrete groups, and relate this to numerical invariants that are computable in terms of characteristic classes. In the spin setting, these results can be applied to give new obstructions to metrics of positive scalar curvature that are invariant under the proper action of a discrete group. This generalises the obstructions provided by the higher A^\widehat{A}-genera of Gromov and Lawson [9] to the setting of proper actions.

On the operator-algebraic side, where the higher index resides, we consider appropriately weighted traces on twisted group CC^{*}-algebras. These traces generalise the traces associated to conjugacy classes in the setting of ordinary group CC^{*}-algebras by taking into account a U(1)\mathrm{U}(1)-valued group 22-cocycle. On the geometric side, there is a corresponding algebra of projectively invariant operators of an appropriate trace class, in which the heat operator associated to a projectively invariant Dirac operator lies. The traces on these two sides are related through a weighted L2L^{2}-Lefschetz fixed-point theorem that generalises a result of Wang-Wang [24, Theorem 6.1].

The special case where the group Γ\Gamma is torsion free and was considered in [6, 17], and led to a proof that the higher A^\widehat{A}-genera of MM arising from the subring of H(BΓ,)H^{*}(B\Gamma,\mathbb{R}) generated by elements of degree at most 22 are obstructions to positive scalar curvature on M/ΓM/\Gamma. This gave a partial answer to a conjecture of Gromov-Lawson [9, Conjecture]; see also [22, Conjecture 2.1].

Now let us turn to the general case where an arbitrary discrete group Γ\Gamma, possibly with torsion, acts properly on a manifold MM. Denote by B¯Γ\underline{B}\Gamma the classifying space for proper actions [2], and let

f:M/ΓB¯Γf\colon M/\Gamma\to\underline{B}\Gamma

be the classifying map for the action of Γ\Gamma on MM. For any integer m0m\geq 0, let αZm(B¯Γ,)\alpha\in Z^{m}(\underline{B}\Gamma,\mathbb{R}). Let

ωΩm(M)\omega\in\Omega^{m}(M) (1.1)

be the Γ\Gamma-invariant lift of a differential form on the orbifold M/ΓM/\Gamma belonging to the class f[α]f^{*}[\alpha] in the orbifold de Rham cohomology group HdRm(M/Γ)H^{m}_{\textnormal{dR}}(M/\Gamma). (For background on orbifold de Rham cohomology, see for example [1, chapter 2].)

If the quotient M/ΓM/\Gamma is compact, then for each gΓg\in\Gamma, the centraliser ZgZ^{g} of gg acts properly and cocompactly on the fixed-point submanifold MgM^{g}. Let cgc^{g} be a cut-off function for this action, i.e. cc is non-negative and satisfies

kZgcg(k1x)=1\sum_{k\in Z^{g}}c^{g}(k^{-1}x)=1

for any xMgx\in M^{g}. We define the higher localised A^\widehat{A}-genus of MM with respect to ω\omega to be

A^g(M,ω)MgcgA^(Mg)ω|Mgdet(1geR𝒩/2πi)1/2,\widehat{A}_{g}(M,\omega)\coloneqq\int_{M^{g}}c^{g}\cdot\frac{\widehat{A}(M^{g})\cdot\omega|_{M^{g}}}{\det(1-ge^{-R^{\mathcal{N}}/2\pi i})^{1/2}}, (1.2)

where R𝒩R^{\mathcal{N}} is the curvature of the Levi–Civita connection restricted to the normal bundle 𝒩\mathcal{N} of MgM^{g} in MM with respect to a Γ\Gamma-invariant Riemannian metric on MM.

We are led to the following conjecture:

Conjecture 1.1.

If a proper, cocompact, connected Γ\Gamma-spin manifold MM admits a Γ\Gamma-invariant Riemannian metric of positive scalar curvature, then all of the higher localised A^\widehat{A}-genera of MM vanish. That is, if ω\omega is any closed mm-form constructed as above, then for all gΓg\in\Gamma, we have

A^g(M,ω)=0.\widehat{A}_{g}(M,\omega)=0.
Remark 1.2.

We give evidence for the validity of this conjecture in Theorem 1.5 and Corollary 1.6 by proving it for all ω\omega arising from the cohomology ring generated by fH1(B¯Γ,)fH2(B¯Γ,)f^{*}H^{1}(\underline{B}\Gamma,\mathbb{R})\cup f^{*}H^{2}(\underline{B}\Gamma,\mathbb{R}), under the assumption that MgM^{g} is connected and gg is a regular element for the relevant group multiplier on Γ\Gamma (see subsection 2.2). In view of Remark 4.11, it seems likely that, in the special case of such ω\omega and regular gg, Conjecture 1.1 is implied by the Baum-Connes conjecture, without any growth conditions on (g)(g).

Remark 1.3.

Conjecture 1.1 is a special case of a possible generalisation of the Gromov-Lawson-Rosenberg conjecture in KOKO-theory [22]. We will not discuss this here, since our results are not in this direction.

The following is an immediate corollary to Conjecture 1.1:

Corollary 1.4.

Suppose E¯Γ\underline{E}\Gamma is a proper, cocompact Γ\Gamma-spin manifold. Then there is no Γ\Gamma-invariant metric of positive scalar curvature on E¯Γ\underline{E}\Gamma.

Indeed, suppose such a Riemannian metric existed. Letting ω\omega be the associated volume form and g=eg=e, the quantity (1.2) is A^e(E¯Γ,ω)=Mcω\widehat{A}_{e}(\underline{E}\Gamma,\omega)=\int_{M}c\omega, where cc is a cut-off function for the Γ\Gamma-action on MM. This is positive, which contradicts Conjecture 1.1.

We give an index-theoretic approach to a special case of Conjecture 1.1 as follows. From the data above, we construct a twisted Dirac operator that is invariant under a projective representation of Γ\Gamma. We show that the integral (1.2) arises naturally as certain weighted traces of the associated heat operator, and relate this to the higher index of the twisted Dirac operator via a weighted trace map on the twisted group algebra. We prove:

Theorem 1.5.

Let Γ\Gamma be a finitely generated group, and let MM be a connected, Γ\Gamma-equivariantly spin manifold such that M/ΓM/\Gamma is compact and H1(M)=0H^{1}(M)=0. Let ωΩ2(M)\omega\in\Omega^{2}(M) be as in (2.1), and let DD be the associated projectively invariant Dirac operator on MM from Definition 2.4. Let α\alpha and σ\sigma be the multipliers constructed in subsection 2.2, and suppose that gΓg\in\Gamma is an α\alpha-regular element, in the sense of (3.4), whose conjugacy class (g)(g) has polynomial growth.

  1. (i)

    Suppose MM is even-dimensional. Then

    (τσ(g))IndΓ,σ(D)=j=1mMjgeiϕg(xj)cgA^(Mjg)eω/2πi|Mjgdet(1geR𝒩j/2πi)1/2\big{(}\tau^{(g)}_{\sigma}\big{)}_{*}\operatorname{Ind}_{\Gamma,\sigma}(D)=\sum_{j=1}^{m}\int_{M^{g}_{j}}e^{-i\phi_{g}(x_{j})}c^{g}\cdot\frac{\widehat{A}(M^{g}_{j})\cdot e^{-\omega/2\pi i}|_{M^{g}_{j}}}{\det(1-ge^{-R^{\mathcal{N}_{j}}/2\pi i})^{1/2}}

    where IndΓ,σ(D)K0(Cr(Γ,σ))\operatorname{Ind}_{\Gamma,\sigma}(D)\in K_{0}(C^{*}_{r}(\Gamma,\sigma)) be the (Γ,σ)(\Gamma,\sigma)-invariant higher index of DD from Definition 3.32, (τσ(g))\big{(}\tau^{(g)}_{\sigma}\big{)}_{*} is the homomorphism (4.8), M1g,,MmgM^{g}_{1},\ldots,M^{g}_{m} are the connected components of MgM^{g} that intersect the support of a cut-off function cgc^{g} for the action of ZgZ^{g} on MgM^{g}, ϕg\phi_{g} is as in (2.2), xjx_{j} is an arbitrary point in MjgM^{g}_{j}, and R𝒩jR^{\mathcal{N}_{j}} is the curvature of the Levi-Civita connection restricted to the normal bundle 𝒩j\mathcal{N}_{j} of MjgM^{g}_{j} in MM and with respect to a Γ\Gamma-invariant Riemannian metric on MM.

  2. (ii)

    If MM admits a Γ\Gamma-invariant Riemannian metric of positive scalar curvature, then for each integer k0k\geq 0, we have

    j=1mMjgcgA^(Mg)(ϕg(xj)ω2π)k|Mjgdet(1geR𝒩j/2πi)1/2=0,\sum_{j=1}^{m}\int_{M^{g}_{j}}c^{g}\cdot\frac{\widehat{A}(M^{g})\cdot\left(\phi_{g}(x_{j})-\frac{\omega}{2\pi}\right)^{k}|_{M^{g}_{j}}}{\det(1-ge^{-R^{\mathcal{N}_{j}}/2\pi i})^{1/2}}=0, (1.3)

    where the notation is as explained in part (i). If MgM^{g} is connected, then

    A^g(M,ωk)=0\widehat{A}_{g}(M,\omega^{k})=0

    for each k0k\geq 0.

Corollary 1.6.

Let Γ\Gamma, MM, and gg be as in the statement of Theorem 1.5. If MgM^{g} is connected, then Conjecture 1.1 holds for ω\omega arising from the subring of H(B¯Γ,)H^{*}(\underline{B}\Gamma,\mathbb{R}) generated by fH1(B¯Γ,)fH2(B¯Γ,)f^{*}H^{1}(\underline{B}\Gamma,\mathbb{R})\cup f^{*}H^{2}(\underline{B}\Gamma,\mathbb{R}).

Next, we prove two obstruction results when M/ΓM/\Gamma is non-compact, again via index theory of projectively invariant Dirac operators. First, we show that the higher localised A^\widehat{A}-genera are obstructions to Γ\Gamma-invariant Riemannian metrics with positive scalar curvature in a neighbourhood of MM. More precisely, we prove:

Theorem 1.7.

Let MM be a connected spin manifold on which a discrete group Γ\Gamma acts properly, preserving the spin structure, and suppose H1(M)=0H^{1}(M)=0. Let HH be a connected Γ\Gamma-cocompact hypersurface in MM with trivial normal bundle. Let the multipliers α\alpha and σ\sigma be as in subsection 5.2, and suppose that gΓg\in\Gamma is an α\alpha-regular element, in the sense of (3.4), whose conjugacy class (g)(g) has polynomial growth.

Suppose MM admits a complete Γ\Gamma-invariant Riemannian metric whose scalar curvature is non-negative everywhere on MM and positive in a neighbourhood of HH. Let ω\omega be as above in (1.2). Then if HgH^{g} is connected,

A^g(H,ωk|H)=0\widehat{A}_{g}(H,\omega^{k}|_{H})=0

for each integer k0k\geq 0. In particular,

j=1mHjgeiϕg(xj)cgA^(Hjg)eω/2πi|Hjgdet(1geR𝒩j/2πi)1/2=0,\sum_{j=1}^{m}\int_{H^{g}_{j}}e^{-i\phi_{g}(x_{j})}c^{g}\cdot\frac{\widehat{A}(H^{g}_{j})\cdot e^{-\omega/2\pi i}|_{H^{g}_{j}}}{\det(1-ge^{-R^{\mathcal{N}_{j}}/2\pi i})^{1/2}}=0,

where gΓg\in\Gamma is a σ\sigma-regular element, H1g,,HmgH^{g}_{1},\ldots,H^{g}_{m} are the connected components of the fixed-point set HgH^{g} that intersect the support of a cut-off function cHgc^{g}_{H} for the ZgZ^{g}-action on HgH^{g}, and 𝒩j\mathcal{N}_{j} is the normal bundle of HjgH^{g}_{j} in HH.

The proof of this theorem uses a Callias-type index theorem for projectively invariant Dirac operators, which is discussed in section 5.

Remark 1.8.

Instead of the assumption that (g)(g) has polynomial growth, the conclusions of Theorem 1.5 and Theorem 1.7 also hold whenever the trace τσs(g):σsΓ\tau^{(g)}_{\sigma^{s}}\colon\mathbb{C}^{\sigma^{s}}\Gamma\to\mathbb{C} extends continuously to Cr(Γ,σs)C^{*}_{r}(\Gamma,\sigma^{s}) for all ss in a small interval (0,δ)(0,\delta).

Second, we show that the projectively invariant higher index is compatible with the framework of quantitative KK-theory [21], which is a refinement of operator KK-theory in the context of geometric CC^{*}-algebras. As shown in subsection 6.1, this allows us to formulate a quantitative notion of projectively invariant higher index, generalising that defined in [12]. With respect to this index, we obtain a parameterised vanishing theorem where the vanishing propagation depends on the lower bound of the twisting curvature of the projective Dirac operator as well as the scalar curvature. More precisely, we prove:

Theorem 1.9.

Fix 0<ε<1200<\varepsilon<\frac{1}{20} and N7N\geq 7. There exists a constant λ0\lambda_{0} such that the following holds. Let MM be a smooth spin Riemannian manifold equipped with a proper isometric action by a discrete group Γ\Gamma, and suppose that H1(M)=0H^{1}(M)=0. Let p:MM/Γp\colon M\to M/\Gamma be the projection, κ\kappa the scalar curvature on MM, and 𝒮M\mathcal{S}\to M the spinor bundle.

Let ωΩ2(M)\omega\in\Omega^{2}(M) be as in (2.1), and let DsD^{s} be the associated projectively invariant Dirac operator on MM from Definition 2.4, acting on sections of the bundle 𝒮\mathcal{S}_{\mathscr{L}}. Let κ\kappa denote the scalar curvature on MM, and let cc denote Clifford multiplication. If the estimate

κ+4isc(ω)Cs\kappa+4isc(\omega)\geq C_{s}

holds for some positive constant CsC_{s}, then the quantitative (Γ,σs)(\Gamma,\sigma^{s})-equivariant higher index of DsD^{s} on L2(𝒮)L^{2}(\mathcal{S}_{\mathscr{L}}) at scale rr vanishes for all rλ0Csr\geq\frac{\lambda_{0}}{\sqrt{C_{s}}}:

IndΓ,σs,L2ε,r,N(Ds)=0Kε,r,N(C(M;L2(𝒮))Γ,σs),\operatorname{Ind}_{\Gamma,\sigma^{s},L^{2}}^{\varepsilon,r,N}(D^{s})=0\in K_{*}^{\varepsilon,r,N}(C^{*}(M;L^{2}(\mathcal{S}_{\mathscr{L}}))^{\Gamma,\sigma^{s}}),

where the algebra C(M;L2(𝒮))Γ,σsC^{*}(M;L^{2}(\mathcal{S}_{\mathscr{L}}))^{\Gamma,\sigma^{s}} is as in Definition 3.27, and IndΓ,σs,L2ε,r,N(Ds)\operatorname{Ind}_{\Gamma,\sigma^{s},L^{2}}^{\varepsilon,r,N}(D^{s}) is defined as in subsection 6.2.

Remark 1.10.

When s=0s=0, Theorem 1.9 reduces to [12, Theorem 1.1].

The paper is organised as follows. In section 2, we formulate the preliminary definitions and properties of the relevant operator algebras and projectively invariant operators. In section 4, we prove Theorem 1.5 and Corollary 1.6. In section 5, we develop some Callias-type index theory in the projective setting and use it to prove Theorem 1.7. In section 6, we formulate the quantitative twisted higher index and prove Theorem 1.9.


2. Preliminaries

We first fix some notation and recall the necessary operator-algebraic and geometric terminology we will need.

2.1. Notation

For XX a Riemannian manifold, we write B(X)B(X), Cb(X)C_{b}(X), C0(X)C_{0}(X), and Cc(X)C_{c}(X) to denote the CC^{*}-algebras of complex-valued functions on XX that are, respectively: bounded Borel, bounded continuous, continuous and vanishing at infinity, and continuous with compact support. If SXS\subseteq X is a Borel subset, we write 𝟙S\mathbbm{1}_{S} for the associated characteristic function.

For any CC^{*}-algebra AA, denote its unitization by A+A^{+}. If \mathcal{E} is a Hilbert module over AA, let (){\mathcal{B}}(\mathcal{E}) and 𝒦(){\mathcal{K}}(\mathcal{E}) denote the CC^{*}-algebras of bounded adjointable and compact operators on \mathcal{E} respectively.

For an element gg of a group GG, we use ZgZ^{g} to denote the centraliser of gg in GG.

2.2. Multipliers and projective representations

Let Γ\Gamma be a discrete group.

Definition 2.1.

A multiplier on Γ\Gamma is a map σ:Γ×ΓU(1)\sigma\colon\Gamma\times\Gamma\to\mathrm{U}(1) satisfying

  1. (i)

    σ(γ,μ)σ(γμ,δ)=σ(γ,μδ)σ(μ,δ)\sigma(\gamma,\mu)\sigma(\gamma\mu,\delta)=\sigma(\gamma,\mu\delta)\sigma(\mu,\delta);

  2. (ii)

    σ(γ,γ1)=σ(γ1,γ)=1\sigma(\gamma,\gamma^{-1})=\sigma(\gamma^{-1},\gamma)=1,

for all γ,μ,δΓ\gamma,\mu,\delta\in\Gamma, where ee is the identity element in Γ\Gamma.

In other words, a multiplier on Γ\Gamma is an element of Z2(Γ,U(1))Z^{2}(\Gamma,\operatorname{U}(1)), i.e. a U(1)U(1)-valued group 22-cocycle, satisfying the additional normalisation condition (ii). This condition is slightly stronger than the normalisation requirement in [17], namely σ(e,γ)=σ(γ,e)=1\sigma(e,\gamma)=\sigma(\gamma,e)=1, and we have adopted it to simplify some calculations. Nevertheless, every multiplier in the sense of [17] is cohomologous to a multiplier in our sense. Observe that given a multiplier σ\sigma, its pointwise complex conjugate σ¯\bar{\sigma} is also a multiplier.

We will be concerned specifically with multipliers that arise from proper Γ\Gamma-actions on manifolds in the following way. Let MM be a smooth, connected manifold equipped with a proper action by a discrete group Γ\Gamma, preserving the spin structure. Suppose M/ΓM/\Gamma is compact and that H1(M)=0H^{1}(M)=0. Let B¯Γ\underline{B}\Gamma be the classifying space for proper Γ\Gamma-actions [2], and f:M/ΓB¯Γf\colon M/\Gamma\to\underline{B}\Gamma the classifying map for MM.

Let [β]H2(B¯Γ,)[\beta]\in H^{2}(\underline{B}\Gamma,\mathbb{R}) be a 22-cocycle on B¯Γ\underline{B}\Gamma, and ω0\omega_{0} a closed 22-form on MM representing the de Rham cohomology class f[β]HdR2(M/Γ)f^{*}[\beta]\in H^{2}_{\textnormal{dR}}(M/\Gamma) on the orbifold M/ΓM/\Gamma. Since [β][\beta] lifts trivially to E¯Γ\underline{E}\Gamma, the lift ω\omega of ω0\omega_{0} to MM is exact, so there exists a one-form η\eta (not necessarily Γ\Gamma-invariant) such that

ω=dη.\omega=d\eta. (2.1)

Since ω\omega is Γ\Gamma-invariant, d(γηη)=γωω=0d(\gamma^{*}\eta-\eta)=\gamma^{*}\omega-\omega=0 for all γΓ\gamma\in\Gamma. Thus γηη\gamma^{*}\eta-\eta is a closed 11-form on MM, and hence exact by assumption. It follows that there exists a family

ϕ{ϕγ:γΓ}\phi\coloneqq\{\phi_{\gamma}\colon\gamma\in\Gamma\}

of smooth functions on MM such that

γηη=dϕγ.\gamma^{*}\eta-\eta=d\phi_{\gamma}. (2.2)

This implies that for any γ,γΓ\gamma,\gamma^{\prime}\in\Gamma,

d(ϕγ+γ1ϕγϕγγ)=0.d(\phi_{\gamma}+\gamma^{-1}\phi_{\gamma^{\prime}}-\phi_{\gamma^{\prime}\gamma})=0. (2.3)

By way of normalisation, we may assume that there exists some x0x_{0} such that

ϕγ(γ1x0)=0\phi_{\gamma}(\gamma^{-1}x_{0})=0 (2.4)

for each γΓ\gamma\in\Gamma. This, together with (2.3), implies that ϕe0\phi_{e}\equiv 0 and that the formula

αϕ(γ,γ)=12π(ϕγ(x0)+ϕγ(γx0)ϕγγ(x0))\alpha_{\phi}(\gamma,\gamma^{\prime})=\frac{1}{2\pi}(\phi_{\gamma^{\prime}}(x_{0})+\phi_{\gamma}(\gamma^{\prime}x_{0})-\phi_{\gamma\gamma^{\prime}}(x_{0})) (2.5)

defines an element of Z2(Γ,)Z^{2}(\Gamma,\mathbb{R}). For each ss\in\mathbb{R}, we have an associated U(1)\operatorname{U}(1)-valued 22-cocycle

σϕs(γ,γ)e2πisαϕ(γ,γ),\sigma^{s}_{\phi}(\gamma,\gamma^{\prime})\coloneqq e^{2\pi is\alpha_{\phi}(\gamma,\gamma^{\prime})}, (2.6)

which is a multiplier the sense of Definition 2.1. When it is clear from context, we will use the following abbreviations:

α=αϕ,σs=σϕs,σ=σ1.\alpha=\alpha_{\phi},\qquad\sigma^{s}=\sigma_{\phi}^{s},\qquad\sigma=\sigma^{1}.

It can be shown that for a given class [β][\beta], different choices of ω\omega, η\eta, and ϕ\phi all lead to cohomologous σ\sigma. Nevertheless, it is useful to make specific choices, as the numerical obstructions we compute in Theorems 1.5 and 1.7 are expressed in terms of ϕ\phi.

Remark 2.2.

Restricting (2.2) to the fixed-point submanifold MγM^{\gamma}, one sees that

γη|Mγη|Mγ=dMγϕγ=0,\gamma^{*}\eta|_{M^{\gamma}}-\eta|_{M^{\gamma}}=d_{M^{\gamma}}\phi_{\gamma}=0,

hence on any connected component MγM^{\gamma}, the function ϕγ\phi_{\gamma} is constant.

Definition 2.3.

Let EME\to M be a Γ\Gamma-equivariant \mathbb{C}-vector bundle. For each γΓ\gamma\in\Gamma and ss\in\mathbb{R}, define the unitary operators UγU_{\gamma}, SγsS^{s}_{\gamma}, and TγsT^{s}_{\gamma} on L2(E)L^{2}(E) by:

  • Uγu(x)=γu(γ1x)U_{\gamma}u(x)=\gamma u(\gamma^{-1}x);

  • Sγsu=eisϕγuS_{\gamma}^{s}u=e^{is\phi_{\gamma}}u;

  • Tγs=UγSγsT_{\gamma}^{s}=U_{\gamma}\circ S_{\gamma}^{s},

for uL2(E)u\in L^{2}(E) and xMx\in M. We refer to TsT^{s} as a projective action on L2(E)L^{2}(E).

Note that for any γ,γΓ\gamma,\gamma^{\prime}\in\Gamma and ss\in\mathbb{R}, we have

TγsTγs=σs(γ,γ)Tγγs.T_{\gamma}^{s}T_{\gamma^{\prime}}^{s}=\sigma^{s}(\gamma,\gamma^{\prime})T_{\gamma\gamma^{\prime}}^{s}.

Thus for each ss, the map Ts:ΓU(L2(E))T^{s}\colon\Gamma\to\operatorname{U}(L^{2}(E)) given by Ts(γ)=TγsT^{s}(\gamma)=T^{s}_{\gamma} defines a projective representation of Γ\Gamma in the sense of subsection 2.3 below. An operator on L2(E)L^{2}(E) that commutes with TsT^{s} is said to be (Γ,σs)(\Gamma,\sigma^{s})-invariant or simply projectively invariant if no confusion arises.

Now suppose MM is Γ\Gamma-equivariantly spin, equipped with a Γ\Gamma-invariant Riemannian metric. Let E=𝒮E=\mathcal{S} be the spinor bundle and ∂̸\not{\partial} the spin-Dirac operator. We can obtain a (Γ,σs)(\Gamma,\sigma^{s})-invariant Dirac operator as follows. Let M\mathscr{L}\to M be a Γ\Gamma-equivariantly trivial line bundle. For each sRs\in R, consider the Hermitian connection

,sd+isη\nabla^{\mathscr{L},s}\coloneqq d+is\eta

on \mathscr{L}. Equip 𝒮=𝒮\mathcal{S}_{\mathscr{L}}=\mathcal{S}\otimes\mathscr{L} with the obvious 2\mathbb{Z}_{2}-grading. Then we have:

Definition 2.4.

For each ss\in\mathbb{R}, we will refer to the operator

Ds∂̸,s:L2(𝒮)L2(𝒮)D^{s}\coloneqq\not{\partial}\otimes\nabla^{\mathscr{L},s}\colon L^{2}(\mathcal{S}_{\mathscr{L}})\to L^{2}(\mathcal{S}_{\mathscr{L}}) (2.7)

as the (Γ,σs)(\Gamma,\sigma^{s})-invariant Dirac operator, or simply a projectively invariant Dirac operator if no confusion arises. We write D=D1D=D^{1}.

One can verify, as in [18, Lemma 3.1], that DsD^{s} commutes with the projective (Γ,σs)(\Gamma,\sigma^{s})-action defined by applying Definition 2.3 to E=𝒮E=\mathcal{S}_{\mathscr{L}}.

2.3. Twisted group CC^{*}-algebras

Given a multiplier σ\sigma on a discrete group Γ\Gamma, a unitary (Γ,σ)(\Gamma,\sigma)-representation, or projective representation, of Γ\Gamma is a map

T:ΓU(H),γTγ,T\colon\Gamma\to\mathrm{U}(H),\qquad\gamma\mapsto T_{\gamma},

for some Hilbert space HH, such that

Te\displaystyle T_{e} =IdH;\displaystyle=\textnormal{Id}_{H};
Tγ1Tγ2\displaystyle T_{\gamma_{1}}\circ T_{\gamma_{2}} =σ(γ1,γ2)Tγ1γ2\displaystyle=\sigma(\gamma_{1},\gamma_{2})T_{\gamma_{1}\gamma_{2}}

for all γ1,γ2Γ\gamma_{1},\gamma_{2}\in\Gamma.

The twisted group algebra σΓ\mathbb{C}^{\sigma}\Gamma is the associative *-algebra over \mathbb{C} with basis {γ¯:γΓ}\{\bar{\gamma}\colon\gamma\in\Gamma\}, where the multiplication and *-operation are given on basis elements by

γ¯1γ¯2=σ(γ1,γ2)γ1γ2¯,γ¯=γ1¯,\bar{\gamma}_{1}\cdot\bar{\gamma}_{2}=\sigma(\gamma_{1},\gamma_{2})\overline{\gamma_{1}\gamma_{2}},\qquad\bar{\gamma}^{*}=\overline{\gamma^{-1}},

and extended linearly and conjugate-linearly respectively. It follows from condition (ii) in Definition 2.1 that the identity element of σΓ\mathbb{C}^{\sigma}\Gamma is e¯\bar{e} and that γ¯1=γ¯\bar{\gamma}^{-1}=\bar{\gamma}^{*}.

The reduced twisted group CC^{*}-algebra is constructed as follows. Consider l2(Γ)l^{2}(\Gamma) with its usual basis {δγ}γΓ\{\delta_{\gamma}\}_{\gamma\in\Gamma}, and define a projective representation

λ:ΓU(l2(Γ))\lambda\colon\Gamma\to\operatorname{U}(l^{2}(\Gamma))

by the following action on basis elements:

λγ1δγ2σ(γ1,γ11γ2)δγ1γ2.\lambda_{\gamma_{1}}\delta_{\gamma_{2}}\coloneqq\sigma(\gamma_{1},\gamma_{1}^{-1}\gamma_{2})\delta_{\gamma_{1}\gamma_{2}}.

Then λ\lambda extends naturally to an injective *-representation on (l2(Γ))\mathcal{B}(l^{2}(\Gamma)), called the left regular representation of σΓ\mathbb{C}^{\sigma}\Gamma. The reduced twisted group CC^{*}-algebra Cr(Γ,σ)C^{*}_{r}(\Gamma,\sigma) is the completion of σΓ\mathbb{C}^{\sigma}\Gamma with respect to the induced norm. When convenient, we will simply abbreviate λγ\lambda_{\gamma} to γ¯\bar{\gamma}.

When σ1\sigma\equiv 1 is the trivial multiplier, σΓ\mathbb{C}^{\sigma}\Gamma and Cr(Γ,σ)C^{*}_{r}(\Gamma,\sigma) are the ordinary group algebra and reduced group CC^{*}-algebra respectively.

Remark 2.5.

More generally, twisted group algebras can be formed using an arbitrary group 22-cocycle instead of a multipler. In this case, we would have γ¯=γ¯1=σ¯(γ,γ1)γ1¯\bar{\gamma}^{*}=\bar{\gamma}^{-1}=\bar{\sigma}(\gamma,\gamma^{-1})\overline{\gamma^{-1}}. We will use this extra flexibility in Proposition 3.9 below.

3. Weighted traces and projective indices

We now define the traces associated to conjugacy classes, in an appropriate sense, on the algebraic part of the twisted group algebra that we will use in this paper.

When σ1\sigma\equiv 1 is the trival multiplier, we recover from σΓ\mathbb{C}^{\sigma}\Gamma the group algebra Γ\mathbb{C}\Gamma. In this case, for any conjugacy class (g)Γ(g)\subseteq\Gamma, the map

τ(g):Γ\displaystyle\tau^{(g)}\colon\mathbb{C}\Gamma \displaystyle\to\mathbb{C}
γΓaγγ\displaystyle\sum_{\gamma\in\Gamma}a_{\gamma}\gamma γ(g)aγ\displaystyle\mapsto\sum_{\gamma\in(g)}a_{\gamma} (3.1)

is a trace. When σ\sigma is non-trivial, this formula ceases in general to define a trace on σΓ\mathbb{C}^{\sigma}\Gamma. However, for conjugacy classes of certain σ\sigma-regular elements, one can define a trace via a weighted sum determined by σ\sigma.

Definition 3.1.

Let σ\sigma be a multiplier on Γ\Gamma. An element gΓg\in\Gamma is σ\sigma-regular if

σ(g,z)=σ(z,g)\sigma(g,z)=\sigma(z,g)

for all zZgz\in Z^{g}.

The following equivalent formulation is useful:

Lemma 3.2.

An element gΓg\in\Gamma is σ\sigma-regular if and only if z¯1g¯z¯=g¯\bar{z}^{-1}\bar{g}\bar{z}=\bar{g} for any zZgz\in Z^{g}.

Proof.

Note that

z¯1g¯z¯=σ(z1,gz)σ(g,z)z1gz¯=σ(z1,zg)σ(g,z)g¯,\bar{z}^{-1}\bar{g}\bar{z}=\sigma(z^{-1},gz)\sigma(g,z)\overline{z^{-1}gz}=\sigma(z^{-1},zg)\sigma(g,z)\bar{g},

where we have used that zg=gzzg=gz, while applying Definition 2.1 shows that

g¯=σ(z1,z)σ(z1z,g)g¯=σ(z1,zg)σ(z,g)g¯.\bar{g}=\sigma(z^{-1},z)\sigma(z^{-1}z,g)\bar{g}=\sigma(z^{-1},zg)\sigma(z,g)\bar{g}.

The two expressions on the right are equal if and only if gg is σ\sigma-regular. ∎

The property of being σ\sigma-regular is invariant under conjugation in Γ\Gamma, hence we may speak of σ\sigma-regular conjugacy classes. The next lemma implies that in order to consider any traces at all on σΓ\mathbb{C}^{\sigma}\Gamma, it is necessary to deal with σ\sigma-regular elements (see also [20, Lemma 1.2]):

Lemma 3.3.

If gg is not σ\sigma-regular, then for any trace map t:σΓt\colon\mathbb{C}^{\sigma}\Gamma\to\mathbb{C}, t(g¯)=0t(\overline{g})=0.

Proof.

If gg is not σ\sigma-regular, then by Lemma 3.2 there exists zZgz\in Z^{g} such that z¯1g¯z¯=λg¯\bar{z}^{-1}\bar{g}\bar{z}=\lambda\bar{g} for λ1\lambda\neq 1. Now

z¯(z¯1g¯)(z¯1g¯)z¯=g¯z¯1g¯z¯=(1λ)g¯.\bar{z}(\bar{z}^{-1}\bar{g})-(\bar{z}^{-1}\bar{g})\bar{z}=\bar{g}-\bar{z}^{-1}\bar{g}\bar{z}=(1-\lambda)\bar{g}.

Since tt is a trace, t(1λ)g¯=0t(1-\lambda)\bar{g}=0 and hence t(g¯)=0t(\bar{g})=0. ∎

To define traces for σ\sigma-regular conjugacy classes, we will use the following weighting function. Define a function θ:ΓU(1)\theta\colon\Gamma\to\operatorname{U}(1) by

θ(γ)={σ(g1,k)σ(k1,g1k)if γ=k1gk for some kΓ,1otherwise.\theta(\gamma)=\begin{cases}\sigma(g^{-1},k)\sigma(k^{-1},g^{-1}k)&\textnormal{if }\gamma=k^{-1}gk\textnormal{ for some $k\in\Gamma$},\\ 1&\textnormal{otherwise.}\end{cases} (3.2)

To prove that θ\theta is well-defined, as well as for later calculations, we will make use of the following lemma.

Lemma 3.4.

For any multiplier σ\sigma on Γ\Gamma and h,kΓh,k\in\Gamma, we have

σ(h,k)σ(k1,h1)=1.\sigma(h,k)\sigma(k^{-1},h^{-1})=1.
Proof.

Note that

e¯\displaystyle\bar{e} =k¯h¯h¯1k¯1=σ(k,h)kh¯h1¯k1¯.\displaystyle=\bar{k}\bar{h}\bar{h}^{-1}\bar{k}^{-1}=\sigma(k,h)\overline{kh}\cdot\overline{h^{-1}}\cdot\overline{k^{-1}}.

Multiplying and simplifying, this equals

σ(k,h)σ(h1,k1)e¯.\sigma(k,h)\sigma(h^{-1},k^{-1})\bar{e}.

The claim follows by equating coefficients of e¯\bar{e}. ∎

Proposition 3.5.

The function θ\theta in (3.2) is well-defined.

Proof.

First note that if γ=k1a1gak\gamma=k^{-1}a^{-1}gak for some aZga\in Z^{g}, then

σ(k1a1,g)σ(k1a1g,ak)k1gk¯\displaystyle\sigma(k^{-1}a^{-1},g)\sigma(k^{-1}a^{-1}g,ak)\overline{k^{-1}gk} =σ(k1a1,g)σ(k1a1g,ak)k1a1gak¯\displaystyle=\sigma(k^{-1}a^{-1},g)\sigma(k^{-1}a^{-1}g,ak)\overline{k^{-1}a^{-1}gak}
=k1a1¯g¯ak¯\displaystyle=\overline{k^{-1}a^{-1}}\cdot\overline{g}\cdot\overline{ak}
=σ¯(k1,a1)σ¯(a,k)k1¯a¯1g¯a¯k¯.\displaystyle=\bar{\sigma}(k^{-1},a^{-1})\bar{\sigma}(a,k)\overline{k^{-1}}\bar{a}^{-1}\bar{g}\bar{a}\bar{k}.

Since aZga\in Z^{g} and gg is σ\sigma-regular, this equals

σ¯(k1,a1)σ¯(a,k)k1¯g¯k¯=σ¯(k1,a1)σ¯(a,k)σ(k1,g)σ(k1g,k)k1gk¯.\displaystyle\bar{\sigma}(k^{-1},a^{-1})\bar{\sigma}(a,k)\overline{k^{-1}}\bar{g}\bar{k}=\bar{\sigma}(k^{-1},a^{-1})\bar{\sigma}(a,k)\sigma(k^{-1},g)\sigma(k^{-1}g,k)\overline{k^{-1}gk}.

By Lemma 3.4, this equals

σ(k1,g)σ(k1g,k)k1gk¯.\sigma(k^{-1},g)\sigma(k^{-1}g,k)\overline{k^{-1}gk}.

Equating coefficients of k1gk¯\overline{k^{-1}gk} shows that θ\theta is well-defined. ∎

Remark 3.6.

Observe that θ(e)=θ(g)=1\theta(e)=\theta(g)=1.

Remark 3.7.

When Γ\Gamma is finite, it is known that the set of σ\sigma-regular conjugacy classes in Γ\Gamma is in bijection with the set of distinct irreducible (Γ,σ)(\Gamma,\sigma)-representations.

Definition 3.8.

Suppose that (g)(g) is σ\sigma-regular for some multiplier σ\sigma on Γ\Gamma. Define the σ\sigma-weighted (g)(g)-trace τσ(g):σΓ\tau^{(g)}_{\sigma}\colon\mathbb{C}^{\sigma}\Gamma\to\mathbb{C} by

γΓaγγ¯γ(g)θ(γ)aγ.\sum_{\gamma\in\Gamma}a_{\gamma}\bar{\gamma}\mapsto\sum_{\gamma\in(g)}\theta(\gamma)a_{\gamma}.

We will show that τσ(g)\tau^{(g)}_{\sigma} is a trace in two steps. Define

σ=σdθ¯,\sigma^{\prime}=\sigma d\bar{\theta},

where dθ¯(γ1,γ2)=θ¯(γ1γ2)θ(γ1)θ(γ2)d\bar{\theta}(\gamma_{1},\gamma_{2})=\bar{\theta}(\gamma_{1}\gamma_{2})\theta(\gamma_{1})\theta(\gamma_{2}) is a 22-coboundary. Then σ\sigma^{\prime} is a 22-cocycle cohomologous to σ\sigma. Let σΓ\mathbb{C}^{\sigma^{\prime}}\Gamma be the twisted group algebra defined using σ\sigma^{\prime} (see Remark 2.5), with basis {γ¯:γΓ}\left\{\underline{\gamma}\colon\gamma\in\Gamma\right\}. We first prove:

Proposition 3.9.

The map τ(g):σΓ\tau^{(g)}\colon\mathbb{C}^{\sigma^{\prime}}\Gamma\to\mathbb{C} defined by

γΓaγγ¯γ(g)aγ\displaystyle\sum_{\gamma\in\Gamma}a_{\gamma}\underline{\gamma}\mapsto\sum_{\gamma\in(g)}a_{\gamma}

is a trace.

Proof.

By definition of the multiplication in σΓ\mathbb{C}^{\sigma^{\prime}}\Gamma, we have for any kΓk\in\Gamma that

k¯1g¯k¯\displaystyle\underline{k}^{-1}\underline{g}\underline{k} =σ¯(k,k1)k1¯g¯k¯\displaystyle=\bar{\sigma}^{\prime}(k,k^{-1})\underline{k^{-1}}\underline{g}\underline{k}
=σ¯(k,k1)σ(g,k)σ(k1,gk)k1gk¯.\displaystyle=\bar{\sigma}^{\prime}(k,k^{-1})\sigma^{\prime}(g,k)\sigma^{\prime}(k^{-1},gk)\underline{k^{-1}gk}.

Using the definition of σ\sigma^{\prime}, this is equal to

σ¯(k,k1)σ(g,k)σ(k1,gk)θ(e)θ¯(k)θ¯(k1)θ¯(gk)θ(g)θ(k)θ¯(k1gk)θ(k1)θ(gk)k1gk¯.\bar{\sigma}(k,k^{-1})\sigma(g,k)\sigma(k^{-1},gk)\theta(e)\bar{\theta}(k)\bar{\theta}(k^{-1})\bar{\theta}(gk)\theta(g)\theta(k)\bar{\theta}(k^{-1}gk)\theta(k^{-1})\theta(gk)\underline{k^{-1}gk}.

Upon cancelling and applying Remark 3.6, together with the definition of θ\theta, this simplifies to k1gk¯\underline{k^{-1}gk}. It follows that

k¯1g¯k¯=k1gk¯\underline{k}^{-1}\underline{g}\underline{k}=\underline{k^{-1}gk} (3.3)

for any kΓk\in\Gamma. By extension, this holds if gg is replaced by any h(g)h\in(g). Indeed, if h=m1gmh=m^{-1}gm, then

k¯1h¯k¯\displaystyle\underline{k}^{-1}\underline{h}\,\underline{k} =k¯1m1gm¯k¯\displaystyle=\underline{k}^{-1}\underline{m^{-1}gm}\underline{k}
=k¯1m¯1g¯m¯k¯\displaystyle=\underline{k}^{-1}\underline{m}^{-1}\underline{g}\underline{m}\,\underline{k}
=σ¯(k,k1)σ¯(m,m1)k1¯m1¯g¯m¯k¯\displaystyle=\bar{\sigma}(k,k^{-1})\bar{\sigma}(m,m^{-1})\underline{k^{-1}}\cdot\underline{m^{-1}}\underline{g}\underline{m}\cdot\underline{k}
=σ¯(k,k1)σ¯(m,m1)σ(m,k)σ(k1,m1)k1¯m1¯g¯m¯k¯\displaystyle=\bar{\sigma}(k,k^{-1})\bar{\sigma}(m,m^{-1})\sigma(m,k)\sigma(k^{-1},m^{-1})\underline{k^{-1}}\cdot\underline{m^{-1}}\underline{g}\underline{m}\cdot\underline{k}
=σ¯(k,k1)σ¯(m,m1)σ(m,k)σ(k1,m1)σ(k1m1,mk)(mk¯)1g¯mk¯.\displaystyle=\bar{\sigma}(k,k^{-1})\bar{\sigma}(m,m^{-1})\sigma(m,k)\sigma(k^{-1},m^{-1})\sigma(k^{-1}m^{-1},mk)(\underline{mk})^{-1}\underline{g}\underline{mk}.

Applying Lemma 3.4 and (3.3), this is equal to

(mk)1gmk¯\displaystyle\underline{(mk)^{-1}gmk} =k1hk¯.\displaystyle=\underline{k^{-1}hk}.

To see that τ(g)\tau^{(g)} is a trace, it suffices to show that for any k,mΓk,m\in\Gamma, we have tr(g)k¯m¯=tr(g)m¯k¯\text{tr}^{(g)}\underline{k}\underline{m}=\text{tr}^{(g)}\underline{m}\underline{k}. We may assume that km=h(g)km=h\in(g), so that k¯m¯=λh¯\underline{k}\underline{m}=\lambda\underline{h} for some λU(1)\lambda\in\operatorname{U}(1). Then clearly tr(g)k¯m¯=λ\text{tr}^{(g)}\underline{k}\underline{m}=\lambda. On the other hand,

m¯k¯=k¯1(k¯m¯)k¯=k¯1(λh¯)k¯=λk¯1h¯k¯.\underline{m}\underline{k}=\underline{k}^{-1}(\underline{k}\underline{m})\underline{k}=\underline{k}^{-1}(\lambda\underline{h})\underline{k}=\lambda\underline{k}^{-1}\underline{h}\,\underline{k}.

By the preceding discussion, this is equal to

λk1hk¯,\lambda\underline{k^{-1}hk},

hence tr(g)m¯k¯=λ\text{tr}^{(g)}\underline{m}\underline{k}=\lambda. ∎

We now use Proposition 3.9 to deduce:

Proposition 3.10.

τσ(g):σΓ\tau^{(g)}_{\sigma}\colon\mathbb{C}^{\sigma}\Gamma\to\mathbb{C} is a trace.

Proof.

Using that σ\sigma^{\prime} is cohomologous to σ\sigma via the coboundary dθ¯d\bar{\theta}, one verifies that the map

fθ:σΓ\displaystyle f_{\theta}\colon\mathbb{C}^{\sigma}\Gamma σΓ,\displaystyle\to\mathbb{C}^{\sigma^{\prime}}\Gamma,
γ¯\displaystyle\bar{\gamma} θ¯(γ)γ¯\displaystyle\mapsto\bar{\theta}(\gamma)\underline{\gamma}

is an isomorphism of *-algebras. We have a commutative diagram

σΓ{\mathbb{C}^{\sigma^{\prime}}\Gamma}{\mathbb{C}}σΓ{\mathbb{C}^{\sigma}\Gamma},{\mathbb{C},}τ(g)\scriptstyle{\tau^{(g)}}fθ\scriptstyle{f_{\theta}}τσ(g)\scriptstyle{\tau_{\sigma}^{(g)}}=\scriptstyle{=}

from which it follows that τσ(g)\tau_{\sigma}^{(g)} is a trace. ∎

For real-valued group cocycles, the notion of regularity also applies: for any αZ2(Γ,)\alpha\in Z^{2}(\Gamma,\mathbb{R}), we say that gΓg\in\Gamma is α\alpha-regular if

α(g,z)=α(z,g)\alpha(g,z)=\alpha(z,g) (3.4)

for all zZgz\in Z^{g}. For such gg, let us define a function ψ:Γ\psi\colon\Gamma\to\mathbb{R} by

ψ(γ)={α(g1,k)+α(k1,g1k)if γ=k1gk for some kΓ,0otherwise.\psi(\gamma)=\begin{cases}\alpha(g^{-1},k)+\alpha(k^{-1},g^{-1}k)&\textnormal{if }\gamma=k^{-1}gk\textnormal{ for some $k\in\Gamma$},\\ 0&\textnormal{otherwise.}\end{cases} (3.5)

As with the function θ\theta in (3.2), one checks that ψ\psi is well-defined.

Corollary 3.11.

Let αZ2(Γ,)\alpha\in Z^{2}(\Gamma,\mathbb{R}) be defined as in (2.5), and let {σs}s\{\sigma^{s}\}_{s\in\mathbb{R}} be the associated family multipliers as in (2.6). If gΓg\in\Gamma is α\alpha-regular, then for every ss\in\mathbb{R},

  1. (i)

    the conjugacy class (g)(g) is σs\sigma^{s}-regular;

  2. (ii)

    the formula

    γΓaγγ¯γ(g)e2πisψ(γ)aγ.\sum_{\gamma\in\Gamma}a_{\gamma}\bar{\gamma}\mapsto\sum_{\gamma\in(g)}e^{2\pi is\psi}(\gamma)a_{\gamma}.

    defines a trace τσs(g):σsΓ\tau^{(g)}_{\sigma^{s}}\colon\mathbb{C}^{\sigma^{s}}\Gamma\to\mathbb{C}, where ψ\psi is as in (3.5).

Proof.

By (2.6) and the fact that gg is α\alpha-regular,

σs(g,z)=e2πisα(g,z)=e2πisα(z,g)=σs(z,g)\sigma^{s}(g,z)=e^{2\pi is\alpha(g,z)}=e^{2\pi is\alpha(z,g)}=\sigma^{s}(z,g)

for any zZgz\in Z^{g}, which proves (i). For (ii), note that the proofs of Propositions 3.9 and 3.10, with θ\theta replaced by

θse2πisψ\theta^{s}\coloneqq e^{2\pi is\psi} (3.6)

and σ\sigma^{\prime} replaced by (σs)=σsdθs(\sigma^{s})^{\prime}=\sigma^{s}d\theta^{s}, imply that τσs(g)\tau^{(g)}_{\sigma^{s}} is a trace for each ss. ∎

3.1. Traces on operators

We now define weighted traces on projectively invariant operators of an appropriate class. Let MM, EE, ϕ\phi, and σ\sigma be as in subsection 2.2.

Definition 3.12.

A continuous function c:M[0,)c\colon M\to[0,\infty) is called a cut-off function for the Γ\Gamma-action on MM if for any xMx\in M we have

gΓc(g1x)=1.\sum_{g\in\Gamma}c(g^{-1}x)=1.
Remark 3.13.

For any proper action, a cut-off function always exists. If the action is cocompact, the cut-off function can be chosen to be compactly supported.

We have the following useful lemma:

Lemma 3.14.

Suppose M/ΓM/\Gamma is compact, and let cc be any cut-off function.

  1. (i)

    Let ff be a smooth function on M/ΓM/\Gamma and f~\widetilde{f} its lift to MM. Then

    Mc(x)f~(x)𝑑x=M/Γf(z)𝑑z.\int_{M}c(x)\widetilde{f}(x)\,dx=\int_{M/\Gamma}f(z)\,dz.
  2. (ii)

    Let hh be a continuous function on M×MM\times M that is invariant under the diagonal action of Γ\Gamma. If c(x)h(x,y),c(y)h(x,y)c(x)h(x,y),c(y)h(x,y) are integrable on M×MM\times M, then

    M×Mc(x)h(x,y)𝑑x𝑑y=M×Mc(y)h(x,y)𝑑x𝑑y.\int_{M\times M}c(x)h(x,y)\,dx\,dy=\int_{M\times M}c(y)h(x,y)\,dx\,dy.
Proof.

Following [24, section 3], let us define the following special cut-off function. First write MM as

M=iΓ×FiUi,M=\bigcup_{i\in\mathbb{N}}\Gamma\times_{F_{i}}U_{i},

for finite subgroups FiF_{i} of Γi\Gamma_{i} and FiF_{i}-invariant, relatively compact subsets UiMU_{i}\subseteq M. Then M/ΓM/\Gamma admits the open cover {Ui/Fi}i\{U_{i}/F_{i}\}_{i\in\mathbb{N}}, along with a subordinate partition of unity {ϕi}i\{\phi_{i}\}_{i\in\mathbb{N}}. Write ϕ~i\widetilde{\phi}_{i} for the Γ\Gamma-invariant lift of ϕi\phi_{i} to MM. For each ii, define the function φi:G×FiUi[0,)\varphi_{i}\colon G\times_{F_{i}}U_{i}\to[0,\infty) by

φi([g,u]){ϕ~i(u)for gFi,0otherwise.\varphi_{i}([g,u])\coloneqq\begin{cases}\widetilde{\phi}_{i}(u)&\textnormal{for }g\in F_{i},\\ 0&\textnormal{otherwise.}\end{cases}

Then φi\varphi_{i} extends by zero to an FiF_{i}-invariant function ψi\psi_{i} on all of MM. Define a smooth function 𝔠:M[0,)\mathfrak{c}\colon M\to[0,\infty) by

𝔠(x)i1|Fi|ψi(x).\mathfrak{c}(x)\coloneqq\sum_{i\in\mathbb{N}}\frac{1}{|F_{i}|}\psi_{i}(x). (3.7)

By [24, Lemma 3.9], 𝔠\mathfrak{c} is a compactly supported, smooth cut-off function on MM whenever the Γ\Gamma-action is cocompact.

The case of c=𝔠c=\mathfrak{c}, where 𝔠\mathfrak{c} is defined as in (3.7), is a special case of [24, Lemma 3.10]. Now observe that for any Γ\Gamma-invariant function rr on MM, the integral Mc(x)r(x)𝑑x\int_{M}c(x)r(x)\,dx is independent of the choice of cut-off function cc. Applying this to r=f~r=\widetilde{f} yields the general case of (i), while letting

r(x)=Mh(x,y)𝑑y,r(y)=Mh(x,y)𝑑xr(x)=\int_{M}h(x,y)\,dy,\qquad r(y)=\int_{M}h(x,y)\,dx

yields the general case of (ii). ∎

Definition 3.15.

Let σ\sigma be a multiplier on Γ\Gamma, and let (g)Γ(g)\subseteq\Gamma be a σ\sigma-regular conjugacy class.

  • A bounded (Γ,σ)(\Gamma,\sigma)-invariant operator SS on L2(𝒮)L^{2}(\mathcal{S}_{\mathscr{L}}) is said to be of (σ,g)(\sigma,g)-trace class if for all ϕ1,ϕ2Cc(M)\phi_{1},\phi_{2}\in C_{c}(M),

    1. (i)

      the operator ϕ1Th1Sϕ2\phi_{1}T_{h^{-1}}S\phi_{2} is of trace class for any h(g)h\in(g);

    2. (ii)

      the sum

      h(g)tr(ϕ1Th1Sϕ2)\sum_{h\in(g)}\operatorname{tr}(\phi_{1}T_{h^{-1}}S\phi_{2}) (3.8)

      converges absolutely.

  • If SS is of (σ,g)(\sigma,g)-trace class, we define its σ\sigma-weighted (g)(g)-trace to be

    trσ(g)(S)h(g)θ(h)tr(c1Th1Sc2),\text{tr}^{(g)}_{\sigma}(S)\coloneqq\sum_{h\in(g)}\theta(h)\cdot\text{tr}(c_{1}T_{h^{-1}}Sc_{2}),

    for some c1,c2Cc(M)c_{1},c_{2}\in C_{c}(M) such that c1c2c_{1}c_{2} is a cut-off function on MM.

Remark 3.16.

If SS has finite propagation, then for any ϕCc(M)\phi\in C_{c}(M), there exists a compactly supported continuous function ff such that ϕS=ϕSf\phi S=\phi Sf. The σ\sigma-weighted (g)(g)-trace of SS can then be equivalently defined to be

h(g)θ(h)tr(cTh1S),\sum_{h\in(g)}\theta(h)\cdot\text{tr}(cT_{h^{-1}}S),

for any cut-off function cc. Compare [24, Definition 3.13] and [25, Definition 3.1].

Lemma 3.17.

If SS is a (Γ,σ)(\Gamma,\sigma)-invariant (σ,g)(\sigma,g)-trace class operator, then

trσ(g)(S)=h(g)θ(h)Meiϕh(x)c(x)Tr(h1KS(hx,x))𝑑x,\textnormal{tr}^{(g)}_{\sigma}(S)=\sum_{h\in(g)}\theta(h)\int_{M}e^{-i\phi_{h}(x)}c(x)\textnormal{Tr}(h^{-1}K_{S}(hx,x))\,dx,

where cc is a cut-off function as in Definition 3.15, and KSK_{S} denotes the Schwartz kernel of SS.

Proof.

Let c1,c2Cc(M)c_{1},c_{2}\in C_{c}(M) such that c1c2=cc_{1}c_{2}=c. Note that for any uL2(𝒮)u\in L^{2}(\mathcal{S}_{\mathscr{L}}) and hΓh\in\Gamma, we have

(Uh1Sc2)u(x)\displaystyle(U_{h}^{-1}Sc_{2})u(x) =Mh1KS(hx,y)c2(y)u(y)𝑑y.\displaystyle=\int_{M}h^{-1}K_{S}(hx,y)c_{2}(y)u(y)\,dy.

Since Th1=Th1=Sh1Uh1T_{h}^{-1}=T_{h^{-1}}=S_{h}^{-1}U_{h}^{-1}, this implies that

(c1Th1Sc2)u(x)=Meiϕh(x)c1(x)h1KS(hx,y)c2(y)u(y)𝑑y.(c_{1}T_{h}^{-1}Sc_{2})u(x)=\int_{M}e^{-i\phi_{h}(x)}c_{1}(x)h^{-1}K_{S}(hx,y)c_{2}(y)u(y)\,dy.

Taking the trace gives

tr(c1Th1Sc2)\displaystyle\textnormal{tr}(c_{1}T_{h^{-1}}Sc_{2}) =Meiϕh(x)c1(x)Tr(h1KS(hx,x))c2(x)𝑑x\displaystyle=\int_{M}e^{-i\phi_{h}(x)}c_{1}(x)\textnormal{Tr}(h^{-1}K_{S}(hx,x))c_{2}(x)\,dx
=Meiϕh(x)c(x)Tr(h1KS(hx,x))𝑑x.\displaystyle=\int_{M}e^{-i\phi_{h}(x)}c(x)\textnormal{Tr}(h^{-1}K_{S}(hx,x))\,dx.

Multiplying by θ(h)\theta(h) and taking a sum finishes the proof. ∎

Our task now is to show that trσ(g)\operatorname{tr}^{(g)}_{\sigma} satisfies the tracial property and that it is independent of the choices of c1c_{1}, c2c_{2}, and cc made in Definition 3.15. This will be carried out in Proposition 3.20, after we record some preparatory observations in the form of Lemmas 3.18 and 3.19.

Lemma 3.18.

Suppose h=x1gxh=x^{-1}gx. Then for any γΓ\gamma\in\Gamma,

σ(γ1h,γ)σ(γ1,h)=σ(γ1x1,g)σ(γ1x1g,xγ)σ¯(x1,g)σ¯(x1g,x).\sigma(\gamma^{-1}h,\gamma)\sigma(\gamma^{-1},h)=\sigma(\gamma^{-1}x^{-1},g)\sigma(\gamma^{-1}x^{-1}g,x\gamma)\bar{\sigma}(x^{-1},g)\bar{\sigma}(x^{-1}g,x).
Proof.

Observe that γ¯1h¯γ¯\bar{\gamma}^{-1}\bar{h}\bar{\gamma} is equal to

γ¯x1gx¯γ¯=γ¯1x¯1g¯x¯γ¯σ¯(x1,g)σ¯(x1g,x).\displaystyle\bar{\gamma}\overline{x^{-1}gx}\bar{\gamma}=\bar{\gamma}^{-1}\bar{x}^{-1}\bar{g}\bar{x}\bar{\gamma}\bar{\sigma}(x^{-1},g)\bar{\sigma}(x^{-1}g,x).

On the other hand, it is also equal to σ(γ1,h)σ(γ1h,γ)γ1hγ¯\sigma(\gamma^{-1},h)\sigma(\gamma^{-1}h,\gamma)\overline{\gamma^{-1}h\gamma}, which, can be written as

σ(γ1,h)σ(γ1h,γ)γ1x1¯g¯xγ¯σ¯(γ1x1,g)σ¯(γ1x1g,xγ).\displaystyle\sigma(\gamma^{-1},h)\sigma(\gamma^{-1}h,\gamma)\overline{\gamma^{-1}x^{-1}}\bar{g}\overline{x\gamma}\bar{\sigma}(\gamma^{-1}x^{-1},g)\bar{\sigma}(\gamma^{-1}x^{-1}g,x\gamma).

By Lemma 3.4, we have γ¯1x¯1g¯x¯γ¯=γ1x1¯g¯xγ¯\bar{\gamma}^{-1}\bar{x}^{-1}\bar{g}\bar{x}\bar{\gamma}=\overline{\gamma^{-1}x^{-1}}\bar{g}\overline{x\gamma}. We conclude by equating coefficients. ∎

Lemma 3.19.

Let (g)Γ(g)\subseteq\Gamma be a σ\sigma-regular conjugacy class for some multiplier σ\sigma. Then for any h(g)h\in(g) and γΓ\gamma\in\Gamma, we have

θ(γ1hγ)σ¯(γ1hγ,γ1)=θ(h)σ¯(γ1,h).\theta(\gamma^{-1}h\gamma)\cdot\bar{\sigma}(\gamma^{-1}h\gamma,\gamma^{-1})=\theta(h)\cdot\bar{\sigma}(\gamma^{-1},h).
Proof.

We have

σ¯(γ1hγ,γ1)\displaystyle\bar{\sigma}(\gamma^{-1}h\gamma,\gamma^{-1}) =σ(γ,γ1h1γ)\displaystyle={\sigma}(\gamma,\gamma^{-1}h^{-1}\gamma)
=σ¯(γ1,h1γ)\displaystyle=\bar{\sigma}(\gamma^{-1},h^{-1}\gamma)
=σ(γ1h,γ).\displaystyle=\sigma(\gamma^{-1}h,\gamma).

The first and third equalities follow from Lemma 3.4; the second follows from the fact that σ(γ,γ1h1γ)σ(γ1,h1γ)=1\sigma(\gamma,\gamma^{-1}h^{-1}\gamma)\sigma(\gamma^{-1},h^{-1}\gamma)=1. Thus it suffices to show that

θ(γ1hγ)θ¯(h)σ(γ1h,γ)σ(γ1,h)=1.\theta(\gamma^{-1}h\gamma)\bar{\theta}(h)\sigma(\gamma^{-1}h,\gamma)\sigma(\gamma^{-1},h)=1. (3.9)

For this, let h=k1gkh=k^{-1}gk. Then by Lemma 3.18,

σ(γ1h,γ)σ(γ1,h)=σ(γ1k1,g)σ(γ1k1g,kγ)σ¯(k1,g)σ¯(k1g,k).\sigma(\gamma^{-1}h,\gamma)\sigma(\gamma^{-1},h)=\sigma(\gamma^{-1}k^{-1},g)\sigma(\gamma^{-1}k^{-1}g,k\gamma)\bar{\sigma}(k^{-1},g)\bar{\sigma}(k^{-1}g,k).

It follows from the definition of θ\theta that the left-hand side of (3.9) equals

σ(g1,kγ)σ(γ1k1,g1kγ)σ¯(g1,k)σ¯(k1,g1k)σ(γ1k1,g)σ(γ1k1g,kγ)σ¯(k1,g)σ¯(k1g,k),\sigma(g^{-1},k\gamma)\sigma(\gamma^{-1}k^{-1},g^{-1}k\gamma)\bar{\sigma}(g^{-1},k)\bar{\sigma}(k^{-1},g^{-1}k)\\ \cdot\sigma(\gamma^{-1}k^{-1},g)\sigma(\gamma^{-1}k^{-1}g,k\gamma)\bar{\sigma}(k^{-1},g)\bar{\sigma}(k^{-1}g,k),

which equals 11 by repeated applications of Lemma 3.4. ∎

With these preparations, we are now ready to prove that trσ(g)\operatorname{tr}^{(g)}_{\sigma} is tracial and well-defined independently of the choice of functions used in Definition 3.15.

Proposition 3.20.

Suppose (g)(g) is a σ\sigma-regular conjugacy class with respect to the multiplier σ\sigma defined by (2.6). Then:

  1. (i)

    trσ(g)\textnormal{tr}^{(g)}_{\sigma} does not depend on the choices of c1c_{1}, c2c_{2}, and cc in Definition 3.15;

  2. (ii)

    if SS and TT are bounded (Γ,σ)(\Gamma,\sigma)-invariant operators on L2(𝒮)L^{2}(\mathcal{S}_{\mathscr{L}}) such that STST and TSTS are (σ,g)(\sigma,g)-trace class, then trσ(g)(ST)=trσ(g)(TS)\textnormal{tr}_{\sigma}^{(g)}(ST)=\textnormal{tr}_{\sigma}^{(g)}(TS).

Proof.

We begin with (i)(i). Let SS be a (σ,g)(\sigma,g)-trace class operator. By the formula in Lemma 3.17, it is clear that for any cut-off function cc, trσ(g)(S)\operatorname{tr}^{(g)}_{\sigma}(S) is independent of the choice of c1c_{1} and c2c_{2} such that c1c2=cc_{1}c_{2}=c. Define the function m1:Mm_{1}\colon M\to\mathbb{C} by

m1(x)h(g)θ(h)eiϕh(x)Tr(h1KS(hx,x)),m_{1}(x)\coloneqq\sum_{h\in(g)}\theta(h)\cdot e^{-i\phi_{h}(x)}\operatorname{Tr}(h^{-1}K_{S}(hx,x)),

so that

trσ(g)(S)\displaystyle\textnormal{tr}^{(g)}_{\sigma}(S) =Mc1(x)m1(x)c2(x)𝑑x,\displaystyle=\int_{M}c_{1}(x)m_{1}(x)c_{2}(x)\,dx,

where c1,c2Cc(M)c_{1},c_{2}\in C_{c}(M) such that c1c2=cc_{1}c_{2}=c for some cut-off function cc. We will show that m1m_{1} is Γ\Gamma-invariant, whence by Lemma 3.14, tr(g)\operatorname{tr}^{(g)} is independent of c.c.

By (Γ,σ)(\Gamma,\sigma)-invariance in Definition 3.28, we see that for any γΓ\gamma\in\Gamma,

m1(γx)\displaystyle m_{1}(\gamma x) =h(g)θ(h)eiϕh(γx)Tr(h1KS(hγx,γx))\displaystyle=\sum_{h\in(g)}\theta(h)\cdot e^{-i\phi_{h}(\gamma x)}\operatorname{Tr}(h^{-1}K_{S}(h\gamma x,\gamma x))
=h(g)θ(h)eiϕh(γx)Tr(h1eiϕγ1(hγx)γKS(γ1hγx,x)γ1eiϕγ1(γx))\displaystyle=\sum_{h\in(g)}\theta(h)\cdot e^{-i\phi_{h}(\gamma x)}\operatorname{Tr}(h^{-1}e^{-i\phi_{\gamma^{-1}}(h\gamma x)}\gamma K_{S}(\gamma^{-1}h\gamma x,x)\gamma^{-1}e^{i\phi_{\gamma^{-1}}(\gamma x)})
=h(g)θ(h)eiϕh(γx)eiϕγ1(hγx)Tr((γ1hγ)1KS(γ1hγx,x)eiϕγ1(γx)).\displaystyle=\sum_{h\in(g)}\theta(h)\cdot e^{-i\phi_{h}(\gamma x)}e^{-i\phi_{\gamma^{-1}}(h\gamma x)}\operatorname{Tr}((\gamma^{-1}h\gamma)^{-1}K_{S}(\gamma^{-1}h\gamma x,x)e^{i\phi_{\gamma^{-1}}(\gamma x)}). (3.10)

We claim that the γ1hγ\gamma^{-1}h\gamma-summand of m1(x)m_{1}(x) equals the hh-summand of m1(γx)m_{1}(\gamma x). To prove this, it suffices to show that

θ(γ1hγ)ei(ϕγ1hγ(x)+ϕγ1(γx))=θ(h)ei(ϕh(γx)+ϕγ1(hγx)).\theta(\gamma^{-1}h\gamma)\cdot e^{-i(\phi_{\gamma^{-1}h\gamma}(x)+\phi_{\gamma^{-1}}(\gamma x))}=\theta(h)\cdot e^{-i(\phi_{h}(\gamma x)+\phi_{\gamma^{-1}}(h\gamma x))}. (3.11)

Applying identity (2.3) with γγ1\gamma\to\gamma^{-1}, γγ1hγ\gamma^{\prime}\to\gamma^{-1}h\gamma, and xγxx\to\gamma x, one sees that the function

xϕγ1hγ(x)+ϕγ1(γx)ϕγ1h(γx)x\mapsto\phi_{\gamma^{-1}h\gamma}(x)+\phi_{\gamma^{-1}}(\gamma x)-\phi_{\gamma^{-1}h}(\gamma x)

is constant on MM. Letting x=x0x=x_{0}, and using the definition of σ\sigma in terms of ϕ\phi given by (2.6) together with (2.4), we see that

θ(γ1hγ)ei(ϕγ1hγ(x0)+ϕγ1(γx0)ϕγ1h(γx0))=θ(γ1hγ)σ¯(γ1,γ)σ(γ1h,γ).\theta(\gamma^{-1}h\gamma)\cdot e^{-i(\phi_{\gamma^{-1}h\gamma}(x_{0})+\phi_{\gamma^{-1}}(\gamma x_{0})-\phi_{\gamma^{-1}h}(\gamma x_{0}))}=\theta(\gamma^{-1}h\gamma)\cdot\bar{\sigma}(\gamma^{-1},\gamma)\sigma(\gamma^{-1}h,\gamma).

We then have

θ(γ1hγ)σ¯(γ1,γ)σ(γ1h,γ)\displaystyle\theta(\gamma^{-1}h\gamma)\cdot\bar{\sigma}(\gamma^{-1},\gamma)\sigma(\gamma^{-1}h,\gamma) =θ(γ1hγ)σ¯(γ1hγ,γ1)\displaystyle=\theta(\gamma^{-1}h\gamma)\cdot\bar{\sigma}(\gamma^{-1}h\gamma,\gamma^{-1})
=θ(h)σ¯(γ1,h)\displaystyle=\theta(h)\bar{\sigma}(\gamma^{-1},h)
=θ(h)σ¯(h,γ)σ¯(γ1,hγ)σ(γ1h,γ)\displaystyle=\theta(h)\bar{\sigma}(h,\gamma)\bar{\sigma}(\gamma^{-1},h\gamma)\sigma(\gamma^{-1}h,\gamma)
=θ(h)ei(ϕh(γx0)+ϕγ1(hγx0)ϕγ1h(γx0)).\displaystyle=\theta(h)\cdot e^{-i(\phi_{h}(\gamma x_{0})+\phi_{\gamma^{-1}}(h\gamma x_{0})-\phi_{\gamma^{-1}h}(\gamma x_{0}))}.

The first equality follows from the cocycle condition. The second follows from Lemma 3.19 and the third is again by the cocycle condition. The final equality follows from (2.6).

Again by (2.3), this equals θ(h)ei(ϕh(γx)+ϕγ1(hγx)ϕγ1h(γx))\theta(h)\cdot e^{-i(\phi_{h}(\gamma x)+\phi_{\gamma^{-1}}(h\gamma x)-\phi_{\gamma^{-1}h}(\gamma x))}. From this (3.11) follows. Summing over elements of (g)(g) then shows that m1m_{1} is Γ\Gamma-invariant, which completes the proof of (i).

For (ii), note that by Lemma 3.17,

trσ(g)(ST)=h(g)θ(h)M×Meiϕh(x)c(x)Tr(h1KS(hx,y)KT(y,x))𝑑x𝑑y,\textnormal{tr}^{(g)}_{\sigma}(ST)=\sum_{h\in(g)}\theta(h)\cdot\int_{M\times M}e^{-i\phi_{h}(x)}c(x)\textnormal{Tr}(h^{-1}K_{S}(hx,y)K_{T}(y,x))\,dx\,dy,

while

trσ(g)(TS)\displaystyle\textnormal{tr}^{(g)}_{\sigma}(TS) =h(g)θ(h)M×Meiϕh(y)c(y)Tr(h1KT(hy,x)KS(x,y))𝑑x𝑑y\displaystyle=\sum_{h\in(g)}\theta(h)\cdot\int_{M\times M}e^{-i\phi_{h}(y)}c(y)\textnormal{Tr}(h^{-1}K_{T}(hy,x)K_{S}(x,y))\,dx\,dy
=h(g)θ(h)M×Meiϕh(h1y)c(y)Tr(h1KT(y,x)KS(x,h1y))𝑑x𝑑y,\displaystyle=\sum_{h\in(g)}\theta(h)\cdot\int_{M\times M}e^{-i\phi_{h}(h^{-1}y)}c(y)\textnormal{Tr}(h^{-1}K_{T}(y,x)K_{S}(x,h^{-1}y))\,dx\,dy, (3.12)

where we have used the change of variable yh1yy\mapsto h^{-1}y and the fact that yc(h1y)y\mapsto c(h^{-1}y) is a cut-off function for any hΓh\in\Gamma.

Now define m2:M×Mm_{2}\colon M\times M\to\mathbb{C} by

m2(x,y)=h(g)θ(h)eiϕh(x)Tr(h1KS(hx,y)KT(y,x))m_{2}(x,y)=\sum_{h\in(g)}\theta(h)\cdot e^{-i\phi_{h}(x)}\textnormal{Tr}(h^{-1}K_{S}(hx,y)K_{T}(y,x))

so that

trσ(g)(ST)=M×Mc(x)m2(x,y)𝑑x𝑑y.\textnormal{tr}^{(g)}_{\sigma}(ST)=\int_{M\times M}c(x)m_{2}(x,y)\,dx\,dy.

We claim that

  1. (1)

    m2m_{2} is Γ\Gamma-equivariant for the diagonal action on M×MM\times M;

  2. (2)

    trσ(g)(TS)=M×Mc(y)m2(x,y)𝑑x𝑑y.\textnormal{tr}^{(g)}_{\sigma}(TS)=\displaystyle\int_{M\times M}c(y)m_{2}(x,y)\,dx\,dy.

To prove claim (1), fix some γΓ\gamma\in\Gamma. By the change of variable hγ1hγh\to\gamma^{-1}h\gamma, we have

m2(x,y)\displaystyle m_{2}(x,y) =γ1hγ(g)θ(γ1hγ)eiϕγ1hγ(x)Tr((γ1hγ)1KS(γ1hγx,y)KT(y,x)).\displaystyle=\sum_{\gamma^{-1}h\gamma\in(g)}\theta(\gamma^{-1}h\gamma)\cdot e^{-i\phi_{\gamma^{-1}h\gamma}(x)}\textnormal{Tr}((\gamma^{-1}h\gamma)^{-1}K_{S}(\gamma^{-1}h\gamma x,y)K_{T}(y,x)).

Since SS and TT are (Γ,σ)(\Gamma,\sigma)-invariant, Lemma 3.29 implies that m2(γx,γy)m_{2}(\gamma x,\gamma y) equals

h(g)θ(h)eiϕh(γx)Tr(h1KS(hγx,γy)KT(γy,γx))\displaystyle\sum_{h\in(g)}\theta(h)\cdot e^{-i\phi_{h}(\gamma x)}\textnormal{Tr}(h^{-1}K_{S}(h\gamma x,\gamma y)K_{T}(\gamma y,\gamma x))
=\displaystyle= h(g)θ(h)eiϕh(γx)Tr(eiϕγ1(hγx)h1γKS(γ1hγx,y)KT(y,x)γ1eiϕγ1(γx))\displaystyle\sum_{h\in(g)}\theta(h)\cdot e^{-i\phi_{h}(\gamma x)}\textnormal{Tr}(e^{-i\phi_{\gamma^{-1}}(h\gamma x)}h^{-1}\gamma K_{S}(\gamma^{-1}h\gamma x,y)K_{T}(y,x)\gamma^{-1}e^{i\phi_{\gamma^{-1}}(\gamma x)})
=\displaystyle= h(g)θ(h)ei(ϕh(γx)+ϕγ1(hγx)ϕγ1(γx))Tr((γ1hγ)1KS(γ1hγx,y)KT(y,x)).\displaystyle\sum_{h\in(g)}\theta(h)\cdot e^{-i(\phi_{h}(\gamma x)+\phi_{\gamma^{-1}}(h\gamma x)-\phi_{\gamma^{-1}}(\gamma x))}\textnormal{Tr}((\gamma^{-1}h\gamma)^{-1}K_{S}(\gamma^{-1}h\gamma x,y)K_{T}(y,x)).

We claim that the γ1hγ\gamma^{-1}h\gamma-summand of m2(x,y)m_{2}(x,y) equals the hh-summand of m2(γx,γy)m_{2}(\gamma x,\gamma y), that is:

θ(γ1hγ)ei(ϕγ1hγ(x))=θ(h)ei(ϕh(γx)+ϕγ1(hγx)ϕγ1(γx));\theta(\gamma^{-1}h\gamma)\cdot e^{-i(\phi_{\gamma^{-1}h\gamma}(x))}=\theta(h)\cdot e^{-i(\phi_{h}(\gamma x)+\phi_{\gamma^{-1}}(h\gamma x)-\phi_{\gamma^{-1}}(\gamma x))}; (3.13)

but this is precisely what we have already established in (3.11). Now summing over elements of (g)(g) proves claim (1).

Now for claim (2), let us denote the hh-summand of m2(x,y)m_{2}(x,y) by [m2(x,y)]h[m_{2}(x,y)]_{h}. Then by (3.1), it suffices to show that for each h(g)h\in(g), we have

[m2(x,y)]h=θ(h)eiϕh(h1y)Tr(h1KT(y,x)KS(x,h1y)).[m_{2}(x,y)]_{h}=\theta(h)\cdot e^{-i\phi_{h}(h^{-1}y)}\textnormal{Tr}(h^{-1}K_{T}(y,x)K_{S}(x,h^{-1}y)). (3.14)

To see this, note that by Lemma 3.29, we have

Tr(h1KS(hx,y)KT(y,x))\displaystyle\textnormal{Tr}(h^{-1}K_{S}(hx,y)K_{T}(y,x)) =Tr(KT(y,x)h1KS(hx,y))\displaystyle=\textnormal{Tr}(K_{T}(y,x)h^{-1}K_{S}(hx,y))
=Tr(KT(y,x)h1eiϕh1(hx)hKS(x,h1y))h1eiϕh1(y))\displaystyle=\textnormal{Tr}(K_{T}(y,x)h^{-1}e^{-i\phi_{h^{-1}}(hx)}hK_{S}(x,h^{-1}y))h^{-1}e^{i\phi_{h^{-1}}(y)})
=Tr(h1KT(y,x)KS(x,h1y))ei(ϕh1(hx)ϕh1(y)).\displaystyle=\textnormal{Tr}(h^{-1}K_{T}(y,x)K_{S}(x,h^{-1}y))e^{-i(\phi_{h^{-1}}(hx)-\phi_{h^{-1}}(y))}.

Using the definition of m2(x,y)m_{2}(x,y), this means that

[m2(x,y)]h=θ(h)ei(ϕh(x)+ϕh1(hx)ϕh1(y))Tr(h1KT(y,x)KS(x,h1y)).[m_{2}(x,y)]_{h}=\theta(h)\cdot e^{-i(\phi_{h}(x)+\phi_{h^{-1}}(hx)-\phi_{h^{-1}}(y))}\textnormal{Tr}(h^{-1}K_{T}(y,x)K_{S}(x,h^{-1}y)).

Thus to establish (3.14), it remains to show that

ϕh(h1y)+ϕh1(y)=ϕh(x)+ϕh1(hx).\phi_{h}(h^{-1}y)+\phi_{h^{-1}}(y)=\phi_{h}(x)+\phi_{h^{-1}}(hx).

By the identity (2.3), the left-hand side equals ϕhh1(y)=ϕe(y)\phi_{hh^{-1}}(y)=\phi_{e}(y), while the right-hand side equals ϕhh1(hx)=ϕe(hx)\phi_{hh^{-1}}(hx)=\phi_{e}(hx). Thus equality follows from the normalisation condition ϕe0\phi_{e}\equiv 0. This completes the proof of claim (2).

Finally, since m2m_{2} is invariant under the diagonal Γ\Gamma-action, we may apply Lemma 3.14 (ii) to m2m_{2}, and use claim (2), to obtain

trσ(g)(ST)\displaystyle\textnormal{tr}^{(g)}_{\sigma}(ST) =M×Mc(x)m2(x,y)𝑑x𝑑y\displaystyle=\int_{M\times M}c(x)m_{2}(x,y)\,dx\,dy
=M×Mc(y)m2(x,y)𝑑x𝑑y\displaystyle=\int_{M\times M}c(y)m_{2}(x,y)\,dx\,dy
=trσ(g)(TS).\displaystyle=\textnormal{tr}^{(g)}_{\sigma}(TS).\qed

To summarise the notation, we now have:

  • the trace τσ(g)\tau^{(g)}_{\sigma} on σΓ\mathbb{C}^{\sigma}\Gamma,

  • the trace Tr\operatorname{Tr} on End(𝒮)\operatorname{End}(\mathcal{S}_{\mathscr{L}});

  • the traces tr\operatorname{tr} and trσ(g)\operatorname{tr}^{(g)}_{\sigma} on operators on L2(𝒮)L^{2}(\mathcal{S}_{\mathscr{L}}).

Recall that for any trace map 𝒯\mathscr{T} on operators on a Hilbert space \mathscr{H}, the associated supertrace applied to an operator

S=(S11S12S21S22)S=\begin{pmatrix}S_{11}&S_{12}\\ S_{21}&S_{22}\end{pmatrix} (3.15)

on the 2\mathbb{Z}_{2}-graded space \mathscr{H}\oplus\mathscr{H} is

𝒯(S11)𝒯(S22).\mathscr{T}(S_{11})-\mathscr{T}(S_{22}).

The supercommutator of homogeneous elements S,T()S,T\in{\mathcal{B}}(\mathscr{H}\oplus\mathscr{H}) is

[S,T]sST(1)degSdegTTS.[S,T]_{s}\coloneqq ST-(-1)^{\textnormal{deg}\,S\cdot\textnormal{deg}\,T}TS.

This extends linearly to arbitrarily elements of ()\mathcal{B}(\mathscr{H}\oplus\mathscr{H}). It follows that the supertrace vanishes on supercommutators.

Definition 3.21.

Denote by Str\operatorname{Str}, str\operatorname{str}, and strσ(g)\operatorname{str}^{(g)}_{\sigma} the supertraces associated to Tr\operatorname{Tr}, tr\operatorname{tr}, and trσ(g)\operatorname{tr}^{(g)}_{\sigma} respectively.

Then Definition 3.15 generalises easily to:

Definition 3.22.

Let SS be a 2\mathbb{Z}_{2}-graded operator on L2(𝒮)L^{2}(\mathcal{S}_{\mathscr{L}}) written in the form (3.15). If S11S_{11} and S22S_{22} are of (σ,g)(\sigma,g)-trace class, then we define the σ\sigma-weighted (g)(g)-supertrace of SS to be

strσ(g)(S)h(g)θ(h)str(c1Th1Sc2).\text{str}_{\sigma}^{(g)}(S)\coloneqq\sum_{h\in(g)}\theta(h)\cdot\text{str}(c_{1}T_{h^{-1}}Sc_{2}). (3.16)
Remark 3.23.

If SS has finite propagation, then (3.16) equals

h(g)θ(h)str(cTh1S),\sum_{h\in(g)}\theta(h)\cdot\text{str}(cT_{h^{-1}}S),

where cc is any cut-off function.

Proposition 3.20 implies:

Corollary 3.24.

If SS and TT are 2\mathbb{Z}_{2}-graded (Γ,σ)(\Gamma,\sigma)-invariant operators on L2(𝒮)L^{2}(\mathcal{S}_{\mathscr{L}}) such that the diagonal entries of STST and TSTS are of (σ,g)(\sigma,g)-trace class, then

strσ(g)[S,T]s=0.\operatorname{str}_{\sigma}^{(g)}[S,T]_{s}=0.
Remark 3.25.

Suppose gg is an α\alpha-regular element, with α\alpha as in (2.5). Then the discussion in this subsection generalises easily to the family of multipliers {σs}s\{\sigma^{s}\}_{s\in\mathbb{R}} from (2.6), once we replace the projective representation TT by TsT^{s} from Definition 2.3, the family ϕ\phi by sϕs\phi, and θ\theta by θs\theta^{s} from (3.6) in the proof of Corollary 3.11.

3.2. Twisted Roe algebras

To formulate the higher index of projectively invariant operators, we will work with certain geometric CC^{*}-algebras. The analogous construction in the non-twisted case is well-known; see for instance [26]. Since the discussion applies to σs\sigma^{s} for any ss, let us fix s=1s=1.

Definition 3.26.

Let AA be an operator on L2(𝒮)L^{2}(\mathcal{S}_{\mathscr{L}}).

  • AA is (Γ,σ)(\Gamma,\sigma)-equivariant if

    TgATg=AT_{g}^{*}AT_{g}=A

    for all gΓg\in\Gamma, where TgT_{g} is as in Definition 2.3;

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

    f1Af2=0;f_{1}Af_{2}=0;
  • The propagation of AA is the extended real number

    prop(A)=sup{d(x,y)|(x,y)supp(A)},\textnormal{prop}(A)=\sup\{d(x,y)\,|\,(x,y)\in\textnormal{supp}(A)\},

    where dd denotes the Riemannian distance on MM;

  • AA is locally compact if fAfA and Af𝒦(L2(𝒮))Af\in\mathcal{K}(L^{2}(\mathcal{S}_{\mathscr{L}})) for all fC0(M)f\in C_{0}(M).

Definition 3.27.

The (Γ,σ)(\Gamma,\sigma)-equivariant algebraic Roe algebra of MM, denoted by [M;L2(𝒮)]Γ,σ\mathbb{C}[M;L^{2}(\mathcal{S}_{\mathscr{L}})]^{\Gamma,\sigma}, is the *-subalgebra of (L2(𝒮))\mathcal{B}(L^{2}(\mathcal{S}_{\mathscr{L}})) consisting of (Γ,σ)(\Gamma,\sigma)-equivariant locally compact operators with finite propagation.

The (Γ,σ)(\Gamma,\sigma)-equivariant Roe algebra of MM, denoted by C(M;L2(𝒮))Γ,σC^{*}(M;L^{2}(\mathcal{S}_{\mathscr{L}}))^{\Gamma,\sigma}, is the completion of [M;L2(𝒮)]Γ,σ\mathbb{C}[M;L^{2}(\mathcal{S}_{\mathscr{L}})]^{\Gamma,\sigma} in (L2(𝒮))\mathcal{B}(L^{2}(\mathcal{S}_{\mathscr{L}})).

Definition 3.28.

Consider the vector bundle Hom(𝒮)=𝒮𝒮M×M\operatorname{Hom}(\mathcal{S}_{\mathscr{L}})=\mathcal{S}_{\mathscr{L}}\boxtimes\mathcal{S}_{\mathscr{L}}\to M\times M. Let 𝒮(M)Γ,σ\mathscr{S}(M)^{\Gamma,\sigma} denote the convolution algebra of smooth sections kk of Hom(𝒮)\operatorname{Hom}(\mathcal{S}_{\mathscr{L}}) such that

  1. (i)

    kk is (Γ,σ)(\Gamma,\sigma)-invariant, in the sense that

    eiϕγ(x)γ1kA(γx,γy)γeiϕγ(y)=kA(x,y)e^{-i\phi_{\gamma}(x)}\gamma^{-1}k_{A}(\gamma x,\gamma y)\gamma e^{i\phi_{\gamma}(y)}=k_{A}(x,y)

    for all γΓ\gamma\in\Gamma and x,yMx,y\in M;

  2. (ii)

    kk has finite propagation, in the sense that there exists an R>0R>0 such that k(x,y)=0k(x,y)=0 whenever d(x,y)>Rd(x,y)>R.

An element k𝒮(M)Γ,σk\in\mathscr{S}(M)^{\Gamma,\sigma} acts on a section uL2(𝒮)u\in L^{2}(\mathcal{S}_{\mathscr{L}}) by

(ku)(x)=Mk(x,y)u(y)𝑑y.(ku)(x)=\int_{M}k(x,y)u(y)\,dy. (3.17)
Lemma 3.29.

Suppose a (Γ,σ)(\Gamma,\sigma)-invariant operator AA on L2(𝒮)L^{2}(\mathcal{S}_{\mathscr{L}}) with finite propagation has a smooth Schwartz kernel kAk_{A}. Then kA𝒮(M)Γ,σk_{A}\in\mathscr{S}(M)^{\Gamma,\sigma}.

Proof.

The finite-propagation property for kAk_{A} follows from the fact that AA has finite propagation. The assumption TγATγ=AT_{\gamma}^{*}AT_{\gamma}=A, together with the fact that Tγ=Tγ1T_{\gamma}^{*}=T_{\gamma}^{-1}, implies that

kA(x,y)\displaystyle k_{A}(x,y) =kTγ1ATγ(x,y)\displaystyle=k_{T_{\gamma}^{-1}AT_{\gamma}}(x,y)
=kSγ1Uγ1AUγSγ(x,y)\displaystyle=k_{S_{\gamma}^{-1}U_{\gamma}^{-1}AU_{\gamma}S_{\gamma}}(x,y)
=eiϕγ(x)kUγ1AUγ(x,y)eiϕγ(y)\displaystyle=e^{-i\phi_{\gamma}(x)}k_{U_{\gamma}^{-1}AU_{\gamma}}(x,y)e^{i\phi_{\gamma}(y)}
=eiϕγ(x)γ1kA(γx,γy)γeiϕγ(y).\displaystyle=e^{-i\phi_{\gamma}(x)}\gamma^{-1}k_{A}(\gamma x,\gamma y)\gamma e^{i\phi_{\gamma}(y)}.\qed

Conversely, an element of 𝒮(M)Γ,σ\mathscr{S}(M)^{\Gamma,\sigma} defines an element of [M;L2(𝒮)]Γ,σ\mathbb{C}[M;L^{2}(\mathcal{S}_{\mathscr{L}})]^{\Gamma,\sigma} via the action (3.17).

Let (L2(𝒮))\mathcal{M}\subseteq{\mathcal{B}}(L^{2}(\mathcal{S}_{\mathscr{L}})) denote the multiplier algebra of the (Γ,σ)(\Gamma,\sigma)-equivariant Roe algebra, and write 𝒬=/C(M;L2(𝒮))Γ,σ\mathcal{Q}=\mathcal{M}/C^{*}(M;L^{2}(\mathcal{S}_{\mathscr{L}}))^{\Gamma,\sigma}. We have a short exact sequence of CC^{*}-algebras

0C(M;L2(𝒮))Γ,σ𝒬0.0\rightarrow C^{*}(M;L^{2}(\mathcal{S}_{\mathscr{L}}))^{\Gamma,\sigma}\rightarrow\mathcal{M}\rightarrow\mathcal{Q}\rightarrow 0. (3.18)

3.3. The twisted higher index

We discuss the index map first in a more general context. Recall that associated to any short exact sequence of CC^{*}-algebras

0IAA/I0,0\rightarrow I\rightarrow A\rightarrow A/I\rightarrow 0,

is a cyclic exact sequence in KK-theory:

K0(I){K_{0}(I)}K0(A){K_{0}(A)}K0(A/I){K_{0}(A/I)}K1(A/I){K_{1}(A/I)}K1(A){K_{1}(A)}K1(I),{K_{1}(I),}1\scriptstyle{\partial_{1}}0\scriptstyle{\partial_{0}}

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

Definition 3.30.
  1. (i)

    0\partial_{0}: let uu be an invertible matrix with entires in A/IA/I representing a class in K1(A/I)K_{1}(A/I). Write

    w=(0u1u0)=(10u1)(1u101)(10u1).w=\begin{pmatrix}0&-u^{-1}\\ u&0\end{pmatrix}=\begin{pmatrix}1&0\\ u&1\end{pmatrix}\begin{pmatrix}1&-u^{-1}\\ 0&1\end{pmatrix}\begin{pmatrix}1&0\\ u&1\end{pmatrix}.

    Then ww lifts to an invertible matrix WW with entries in AA . Then

    P=W(1000)W1P=W\begin{pmatrix}1&0\\ 0&0\end{pmatrix}W^{-1}

    is an idempotent, and we define

    0[u][P][0001]K0(I).\partial_{0}[u]\coloneqq\left[P\right]-\begin{bmatrix}0&0\\ 0&1\end{bmatrix}\in K_{0}(I). (3.19)
  2. (ii)

    1\partial_{1}: let qq be an idempotent matrix with entries in A/IA/I representing a class in K0(A/I)K_{0}(A/I). Let QQ be a lift of qq to a matrix algebra over AA. Then we define

    1[q][e2πiQ]K1(I).\partial_{1}[q]\coloneqq\left[e^{2\pi iQ}\right]\in K_{1}(I). (3.20)

Now let MM be a Riemannian spin manifold on which Γ\Gamma acts properly and isometrically, respecting the spin structure. Let σ\sigma be a multiplier on Γ\Gamma. Let M\mathscr{L}\to M be a trivial line bundle, and let DD be the twisted Dirac operator as in (2.7) acting on smooth sections of 𝒮\mathcal{S}_{\mathscr{L}}. Pick any normalizing function χ:\chi\colon\mathbb{R}\rightarrow\mathbb{R}, i.e. a continuous, odd function such that

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

and form the bounded self-adjoint operator χ(D)\chi(D) on L2(𝒮)L^{2}(\mathcal{S}_{\mathscr{L}}). When dimM\dim M is even, 𝒮\mathcal{S}_{\mathscr{L}} is naturally a direct sum (𝒮+)(𝒮)(\mathcal{S}^{+}\otimes\mathscr{L})\oplus(\mathcal{S}^{-}\otimes\mathscr{L}), and DD and χ(D)\chi(D) are odd-graded:

D=(0DD+0),χ(D)=(0χ(D)χ(D)+0).D=\begin{pmatrix}0&D_{-}\\ D_{+}&0\end{pmatrix},\qquad\chi(D)=\begin{pmatrix}0&\chi(D)_{-}\\ \chi(D)_{+}&0\end{pmatrix}.
Proposition 3.31.

The class of χ(D)\chi(D) in /C(M;L2(𝒮))Γ,σ\mathcal{M}/C^{*}(M;L^{2}(\mathcal{S}_{\mathscr{L}}))^{\Gamma,\sigma} is invertible and independent of the choice of χ\chi, and χ(D)+12\frac{\chi(D)+1}{2} is an idempotent modulo C(M;L2(𝒮))Γ,σC^{*}(M;L^{2}(\mathcal{S}_{\mathscr{L}}))^{\Gamma,\sigma}.

Proof.

Since χ21C0()\chi^{2}-1\in C_{0}(\mathbb{R}), it suffices to show that for any fC0()f\in C_{0}(\mathbb{R}), we have f(D)C(M;L2(𝒮))Γ,σf(D)\in C^{*}(M;L^{2}(\mathcal{S}_{\mathscr{L}}))^{\Gamma,\sigma}. In fact, functions in the Schwartz algebra 𝒮()\mathcal{S}(\mathbb{R}) with compactly supported Fourier transform form a dense subset of C0()C_{0}(\mathbb{R}), we may assume that ff is such a function. In that case, the operator f(D)f(D) is given by a smooth (Γ,σ)(\Gamma,\sigma)-invariant Schwartz kernel, and hence an element of 𝒮(M)Γ,σ\mathscr{S}(M)^{\Gamma,\sigma} (see Definition 3.28). It follows that f(D)C(M;L2(𝒮))Γ,σf(D)\in C^{*}(M;L^{2}(\mathcal{S}_{\mathscr{L}}))^{\Gamma,\sigma}. Finally, since the difference of any two normalizing functions lies in C0()C_{0}(\mathbb{R}), the class of χ(D)\chi(D) does not depend on the choice of normalising function χ\chi. ∎

Applying Definition 3.30 to the short exact sequence (3.18) leads to the following:

Definition 3.32.

For i=0,1i=0,1, let i\partial_{i} be the connecting maps from Definition 3.30. The (Γ,σ)(\Gamma,\sigma)-invariant higher index of DD on L2(𝒮)L^{2}(\mathcal{S}_{\mathscr{L}}) is the element

0[χ(D)+]K0(C(M;L2(𝒮))Γ,σ)\displaystyle\partial_{0}\left[\chi(D)_{+}\right]\in K_{0}\big{(}C^{*}(M;L^{2}(\mathcal{S}_{\mathscr{L}}))^{\Gamma,\sigma}\big{)}\quad if dimM is even,\displaystyle\textnormal{ if $\dim M$ is even},
1[χ(D)+12]K1(C(M;L2(𝒮))Γ,σ)\displaystyle\partial_{1}\left[\tfrac{\chi(D)+1}{2}\right]\in K_{1}\big{(}C^{*}(M;L^{2}(\mathcal{S}_{\mathscr{L}}))^{\Gamma,\sigma}\big{)}\quad if dimM is odd.\displaystyle\textnormal{ if $\dim M$ is odd}.

From this, we can obtain an index in the KK-theory of Cr(Γ,σ)C^{*}_{r}(\Gamma,\sigma) as follows. Define

jσ:Cc(𝒮)\displaystyle j_{\sigma}\colon C_{c}(\mathcal{S}_{\mathscr{L}}) σ[Γ]Cc(𝒮)\displaystyle\hookrightarrow\mathbb{C}^{\sigma}[\Gamma]\otimes C_{c}(\mathcal{S}_{\mathscr{L}})
(jσe)γ\displaystyle(j_{\sigma}e)_{\gamma} =c(Tg1e),\displaystyle=c\cdot(T_{g^{-1}}e), (3.21)

where TT is the (Γ,σ)(\Gamma,\sigma)-representation on L2(𝒮)L^{2}(\mathcal{S}_{\mathscr{L}}). Then jσj_{\sigma} extends to an isometry

jσ:L2(𝒮)l2(Γ)L2(𝒮).j_{\sigma}\colon L^{2}(\mathcal{S}_{\mathscr{L}})\hookrightarrow l^{2}(\Gamma)\otimes L^{2}(\mathcal{S}_{\mathscr{L}}).

The following is easily proved:

Lemma 3.33.

The map jσj_{\sigma} is equivariant with respect to the projective representation TT on L2(𝒮)L^{2}(\mathcal{S}_{\mathscr{L}}) and TL1T^{L}\otimes 1, where TLT^{L} is the left-regular (Γ,σ)(\Gamma,\sigma)-representation on l2(Γ)l^{2}(\Gamma), as in subsection 2.3.

At the level of operators, the map jσj_{\sigma} induces two inclusion maps

 0, 1:(L2(𝒮))\displaystyle\oplus\,0,\,\oplus\,1\colon\mathcal{B}(L^{2}(\mathcal{S}_{\mathscr{L}})) (l2(Γ)L2(𝒮)),\displaystyle\hookrightarrow\mathcal{B}(l^{2}(\Gamma)\otimes L^{2}(\mathcal{S}_{\mathscr{L}})),

defined by first identifying operators on L2(𝒮)L^{2}(\mathcal{S}_{\mathscr{L}}) with operators on j(L2(𝒮))j(L^{2}(\mathcal{S}_{\mathscr{L}})) via conjugation by jj and then extending them by the zero operator and identity operator, respectively, on the orthogonal complement of j(L2(𝒮))j(L^{2}(\mathcal{S}_{\mathscr{L}})). We will write Ti(i)(T)T\oplus i\coloneqq(\oplus\,i)(T), i=0,1i=0,1.

These maps restrict to inclusions

 0:C(M;L2(𝒮))Γ,σ\displaystyle\oplus\,0\colon C^{*}(M;L^{2}(\mathcal{S}_{\mathscr{L}}))^{\Gamma,\sigma} Cr(Γ,σ)𝒦(L2(𝒮)),\displaystyle\to C^{*}_{r}(\Gamma,\sigma)\otimes\mathcal{K}(L^{2}(\mathcal{S}_{\mathscr{L}})),
 1:(C(M;L2(𝒮))Γ,σ)+\displaystyle\oplus\,1\colon(C^{*}(M;L^{2}(\mathcal{S}_{\mathscr{L}}))^{\Gamma,\sigma})^{+} (Cr(Γ,σ)𝒦(L2(𝒮)))+.\displaystyle\to(C^{*}_{r}(\Gamma,\sigma)\otimes\mathcal{K}(L^{2}(\mathcal{S}_{\mathscr{L}})))^{+}.

For each i=0,1i=0,1, the map i\oplus\,i extends in the obvious way to maps between matrix algebras over C(M;L2(𝒮))Γ,σC^{*}(M;L^{2}(\mathcal{S}_{\mathscr{L}}))^{\Gamma,\sigma} and its unitisation, preserving idempotence and invertibility when i=0i=0 and 11 respectively. We get induced maps

(i):Ki(C(M;L2(𝒮))Γ,σ)Ki(Cr(Γ,σ)𝒦(L2(𝒮)))Ki(Cr(Γ,σ)).(\oplus\,i)_{*}\colon K_{i}(C^{*}(M;L^{2}(\mathcal{S}_{\mathscr{L}}))^{\Gamma,\sigma})\to K_{i}(C^{*}_{r}(\Gamma,\sigma)\otimes\mathcal{K}(L^{2}(\mathcal{S}_{\mathscr{L}})))\cong K_{i}(C^{*}_{r}(\Gamma,\sigma)).

This gives an index

IndΓ,σ(D)(i)IndΓ,σ,L2(D)Ki(Cr(Γ,σ)),\operatorname{Ind}_{{\Gamma,\sigma}}(D)\coloneqq(\oplus\,i)_{*}\operatorname{Ind}_{{\Gamma,\sigma,L^{2}}}(D)\in K_{i}(C^{*}_{r}(\Gamma,\sigma)), (3.22)

where i=dimMi=\dim M (mod 22).

Remark 3.34.

When σ1\sigma\equiv 1, (3.22) recovers the usual Γ\Gamma-equivariant higher index of Dirac operators on cocompact manifolds.


4. Projective PSC obstructions on cocompact manifolds

In this section we prove Theorem 1.5. Thus let Γ\Gamma, MM, α\alpha, σ\sigma, and gg be given as in the statement of the theorem.

We begin by describing how the trace τσ(g)\tau^{(g)}_{\sigma} from Definition 3.8 can be extended to a trace on a smooth dense subalgebra of C(Γ,σ)𝒦(L2(𝒮))C^{*}(\Gamma,\sigma)\otimes\mathcal{{\mathcal{K}}}(L^{2}(\mathcal{S}_{\mathscr{L}})). Our construction is based on [8, section 6], [27, subsection 2.2], and [16, section 3], but some extra arguments are needed to make the adaptation to the proper action case, having to do with the form of the embedding jσj_{\sigma} in (3.3).

Let DD and 𝒮\mathcal{S}_{\mathscr{L}} be as in Definition 2.4, with the underlying manifold MM being cocompact with respect to the action of a finitely generated group Γ\Gamma. We will work with the following choice of orthonormal basis of L2(𝒮)L^{2}(\mathcal{S}_{\mathscr{L}}). Let cc be the cut-off function for the Γ\Gamma-action on MM used in the embedding jj from (3.3). Since the support of cc is compact, there exists an orthonormal basis B0={ui}iB_{0}=\{u_{i}\}_{i\in\mathbb{N}} of L2(𝒮)|NL^{2}(\mathcal{S}_{\mathscr{L}})|_{N}, for some relatively compact neighbourhood NN of supp(c)\text{supp}(c) consisting of eigenfunctions of D2|ND^{2}|_{N}. Asymptotically, the eigenvalues λi\lambda_{i} satisfy λiiq\lambda_{i}\sim i^{q} for some constant q>0q>0. Identifying L2(𝒮)|NL^{2}(\mathcal{S}_{\mathscr{L}})|_{N} with l2()l^{2}(\mathbb{N}) via this basis, one checks that a smoothing operator supported in NN lies in the algebra

={(aij)i,j:supi,jikjl|aij|< for all k,l}\mathscr{R}=\{(a_{ij})_{i,j\in\mathbb{N}}\colon\sup_{i,j}i^{k}j^{l}|a_{ij}|<\infty\textnormal{ for all }k,l\in\mathbb{N}\}

of matrices with rapidly decreasing entries [7, chapter 3], which are dense inside 𝒦(l2()){\mathcal{K}}(l^{2}(\mathbb{N})). Complete B0B_{0} to an orthonormal basis B={vi}iB=\{v_{i}\}_{i\in\mathbb{N}} of L2(𝒮)L^{2}(\mathcal{S}_{\mathscr{L}}), and embed l2()l^{2}(\mathbb{N}) inside itself via the map δkδ2k\delta_{k}\mapsto\delta_{2k}. Since this map preserves \mathscr{R}, the previous identification now extends to an identification of L2(𝒮)L^{2}(\mathcal{S}_{\mathscr{L}}) with the new copy of l2()l^{2}(\mathbb{N}) in such a way that for any S𝒮Γ,σS\in\mathscr{S}^{\Gamma,\sigma} (see Definition 3.28), the operator c12Sc12c^{\frac{1}{2}}Sc^{\frac{1}{2}} is an element of \mathscr{R}. Let

I={i:viB0}.I=\{i\in\mathbb{N}\colon v_{i}\in B_{0}\}. (4.1)

Fix a set of generators of Γ\Gamma, and let (γ)=dΓ(γ,e)\ell(\gamma)=d_{\Gamma}(\gamma,e) denote the associated word length function, where dΓd_{\Gamma} is the word metric. Let 𝒟{\mathcal{D}} be the unbounded self-adjoint operator on l2(Γ)l^{2}(\Gamma) given by 𝒟γ¯=(γ)γ¯{\mathcal{D}}\bar{\gamma}=\ell(\gamma)\cdot\bar{\gamma}, where the notation γ¯\bar{\gamma} is as explained in subsection 2.3. Let Δ\Delta be the unbounded self-adjoint operator on l2()l^{2}(\mathbb{N}) given by

Δ(δi)={λiδi if iI,iαδi otherwise,\Delta(\delta_{i})=\begin{cases}\lambda_{i}\delta_{i}&\textnormal{ if }i\in I,\\ i^{\alpha}\delta_{i}&\textnormal{ otherwise,}\end{cases} (4.2)

for ii\in\mathbb{N}, where λi\lambda_{i} is the ithi^{\textnormal{th}} eigenvalue of D2|ND^{2}|_{N}, and the set II is as in (4.1).

Consider also the unbounded operators =[𝒟,]\partial=[{\mathcal{D}},\,\cdot\,] on (l2()){\mathcal{B}}(l^{2}(\mathbb{N})) and ~=[𝒟1,]\widetilde{\partial}=[{\mathcal{D}}\otimes 1,\,\cdot\,] on (l2(Γ)l2()){\mathcal{B}}(l^{2}(\Gamma)\otimes l^{2}(\mathbb{N})). Note that ~\widetilde{\partial} is a closed derivation with domain Dom(~)\textnormal{Dom}(\widetilde{\partial}), which consists of all elements a(l2(Γ)l2())a\in{\mathcal{B}}(l^{2}(\Gamma)\otimes l^{2}(\mathbb{N})) such that aa maps Dom(𝒟1)\textnormal{Dom}({\mathcal{D}}\otimes 1) to itself, and the operator ~(a)=(𝒟1)aa(𝒟1)\widetilde{\partial}(a)=({\mathcal{D}}\otimes 1)\circ a-a\circ({\mathcal{D}}\otimes 1), defined initially on Dom(𝒟1)\textnormal{Dom}({\mathcal{D}}\otimes 1), extends to a bounded operator on l2(Γ)l2()l^{2}(\Gamma)\otimes l^{2}(\mathbb{N}). Define

(Γ,σ)=kDom(~k)(Cr(Γ,σ)𝒦),\mathscr{B}_{\infty}(\Gamma,\sigma)=\bigcap_{k\in\mathbb{N}}\textnormal{Dom}(\widetilde{\partial}^{k})\cap(C^{*}_{r}(\Gamma,\sigma)\otimes\mathcal{K}),

and

(Γ,σ)={a(Γ,σ):~k(a)(1Δ)2 is bounded kN}.\mathscr{B}(\Gamma,\sigma)=\{a\in\mathscr{B}_{\infty}(\Gamma,\sigma)\colon\widetilde{\partial}^{k}(a)\circ(1\otimes\Delta)^{2}\textnormal{ is bounded }\forall k\in N\}. (4.3)

Note that (Γ,σ)\mathscr{B}(\Gamma,\sigma) is a left ideal as well as a Freéchet subalgebra of (Γ,σ)\mathscr{B}_{\infty}(\Gamma,\sigma), with respect to the Fréchet topology given by the family of norms

an=k=01k!~k(a)(1Δ)2,\|a\|_{n}=\sum_{k=0}^{\infty}\frac{1}{k!}\|\widetilde{\partial}^{k}(a)\circ(1\otimes\Delta)^{2}\|,

for nn\in\mathbb{N}, where the norm on the right-hand side is the operator norm.

Define the von Neumann algebra

𝒜(Γ,σ)={a(l2(Γ)l2()):[γ¯1,a]=0γΓ}.\mathscr{A}(\Gamma,\sigma)=\{a\in{\mathcal{B}}(l^{2}(\Gamma)\otimes l^{2}(\mathbb{N}))\colon[\bar{\gamma}\otimes 1,a]=0\,\,\,\forall\gamma\in\Gamma\}. (4.4)

By [16, Lemmas 1.1, 3.2], we have the following useful fact:

Lemma 4.1.

Any a𝒜(Γ,σ)a\in\mathscr{A}(\Gamma,\sigma) can be written as a strongly convergent sum

a=γΓγ¯aγ,a=\sum_{\gamma\in\Gamma}\bar{\gamma}\otimes a_{\gamma},

where aγ(l2())a_{\gamma}\in{\mathcal{B}}(l^{2}(\mathbb{N})) for each γΓ\gamma\in\Gamma. If aγa_{\gamma}\in\mathscr{R} for each γΓ\gamma\in\Gamma and

γ(γ)kaγΔ<,\sum_{\gamma}\ell(\gamma)^{k}\|a_{\gamma}\circ\Delta\|<\infty,

for all k0k\geq 0, then a(Γ,σ)a\in\mathscr{B}(\Gamma,\sigma).

Lemma 4.2.

The *-algebra (Γ,σ)\mathscr{B}(\Gamma,\sigma) is dense in Cr(Γ,σ)𝒦(l2())C^{*}_{r}(\Gamma,\sigma)\otimes\mathcal{K}(l^{2}(\mathbb{N})) and stable under the holomorphic functional calculus.

Proof.

First note that if a=g¯VσΓa=\bar{g}\otimes V\in\mathbb{C}^{\sigma}\Gamma\otimes\mathscr{R}, then

~k(A)(1Δ)2=k(g¯)VΔ2,\widetilde{\partial}^{k}(A)\circ(1\otimes\Delta)^{2}=\partial^{k}(\bar{g})\otimes V\Delta^{2}, (4.5)

which is bounded since both k(g¯)\partial^{k}(\bar{g}) and VΔ2V\Delta^{2} are bounded. It follows that (Γ,σ)\mathscr{B}(\Gamma,\sigma) is dense in Cr(Γ,σ)𝒦(l2())C^{*}_{r}(\Gamma,\sigma)\otimes\mathcal{K}(l^{2}(\mathbb{N})).

Now recall that a subalgebra AA of a Banach algebra BB is called spectral invariant in BB if the invertible elements of the unitisation A+A^{+} are precisely those which are invertible in B+B^{+}. In the case that AA is a Frechet subalgebra of of BB, AA is spectral invariant if and only if AA is stable under the holomorphic functional calculus in BB, by [23]. Further, it is a purely algebraic fact that if II is a left ideal in AA, then II is itself a spectral invariant subalgebra of BB.

By [15, Theorem 1.2], (Γ,σ)\mathscr{B}_{\infty}(\Gamma,\sigma) is dense in Cr(Γ,σ)𝒦(l2())C^{*}_{r}(\Gamma,\sigma)\otimes\mathcal{K}(l^{2}(\mathbb{N})) and closed under the holomorphic functional calculus, and hence a spectral invariant subalgebra of Cr(Γ,σ)𝒦(l2())C^{*}_{r}(\Gamma,\sigma)\otimes\mathcal{K}(l^{2}(\mathbb{N})). Since (Γ,σ)\mathscr{B}(\Gamma,\sigma) is a left ideal in (Γ,σ)\mathscr{B}_{\infty}(\Gamma,\sigma) as well as a Fréchet subalgebra, it follows from the above discussion that (Γ,σ)\mathscr{B}(\Gamma,\sigma) is holomorphically closed. ∎

Let Tr\operatorname{Tr} denote the operator trace on \mathscr{R}, and denote by τσ(g)Tr\tau^{(g)}_{\sigma}\otimes\operatorname{Tr} the amplified trace on Γ\mathbb{C}\Gamma\otimes\mathscr{R}. Recall that a conjugacy class (g)(g) is said to have polynomial growth if there exist constants CC and dd such that the number of elements h(g)h\in(g) such that (h)l\ell(h)\leq l is at most CldCl^{d}.

Lemma 4.3.

Let gΓg\in\Gamma be a σ\sigma-regular element with respect to a multiplier σ\sigma. If the conjugacy class (g)(g) has polynomial growth, then τσ(g)Tr\tau^{(g)}_{\sigma}\otimes\operatorname{Tr} extends to a continuous trace on (Γ,σ)\mathscr{B}(\Gamma,\sigma).

Proof.

This follows from an adaptation of the proof of [27, Lemma 2.7]. Indeed, let 𝒜=(𝒜ij)i,j\mathcal{A}=(\mathcal{A}_{ij})_{i,j\in\mathbb{N}} be an arbitrary element of (Γ,σ)\mathscr{B}(\Gamma,\sigma), where 𝒜ijCr(Γ,σ)\mathcal{A}_{ij}\in C^{*}_{r}(\Gamma,\sigma) can be written as

𝒜ij=γΓ𝒜ij,γγ¯.\mathcal{A}_{ij}=\sum_{\gamma\in\Gamma}\mathcal{A}_{ij,\gamma}\bar{\gamma}. (4.6)

Define the function t:(Γ,σ)t\colon\mathscr{B}(\Gamma,\sigma)\to\mathbb{C} by

t(𝒜)=ih(g)θ(h)𝒜ii,h.t(\mathcal{A})=\sum_{i\in\mathbb{N}}\sum_{h\in(g)}\theta(h)\mathcal{A}_{ii,h}. (4.7)

It follows from Definition 3.8 that tt agrees with τσ(g)Tr\tau^{(g)}_{\sigma}\otimes\operatorname{Tr} on Γ\mathbb{C}\Gamma\otimes\mathscr{R}. Now since θ(γ)U(1)\theta(\gamma)\in\operatorname{U}(1), the right-hand side of (4.7) converges absolutely by the same estimates as in the proof of [27, Lemma 2.7], hence tt extends to a continuous trace on (Γ,σ)\mathscr{B}(\Gamma,\sigma). ∎

We will continue to write τσ(g)Tr\tau^{(g)}_{\sigma}\otimes\operatorname{Tr} for the extended trace tt on (Γ,σ)\mathscr{B}(\Gamma,\sigma) from the proof of Lemma 4.3. It follows from and Lemmas 4.2 and 4.3 that we have an induced map

(τσ(g)):K0(Cr(Γ,σ)𝒦(L2(𝒮)))K0((Γ,σ)).(\tau^{(g)}_{\sigma})_{*}\colon K_{0}(C^{*}_{r}(\Gamma,\sigma)\otimes{\mathcal{K}}(L^{2}(\mathcal{S}_{\mathscr{L}})))\cong K_{0}(\mathscr{B}(\Gamma,\sigma))\to\mathbb{C}. (4.8)

We are now ready to establish the connection between τσ(g)\tau^{(g)}_{\sigma} and the trace trσ(g)\operatorname{tr}_{\sigma}^{(g)} from section 3.

Proposition 4.4.

Let AA be a (Γ,σ)(\Gamma,\sigma)-invariant operator on L2(𝒮)L^{2}(\mathcal{S}_{\mathscr{L}}) and gΓg\in\Gamma a σ\sigma-regular element with respect to σ\sigma. If AA is of (σ,g)(\sigma,g)-trace class and A0(Γ,σ)A\oplus 0\in\mathscr{B}(\Gamma,\sigma), then

(τσ(g)Tr)(A0)=trσ(g)(A).(\tau^{(g)}_{\sigma}\otimes\operatorname{Tr})(A\oplus 0)=\operatorname{tr}^{(g)}_{\sigma}(A).
Proof.

For any γΓ\gamma\in\Gamma, denote by

qγ:l2(Γ)L2(𝒮)δγL2(𝒮)q_{\gamma}\colon l^{2}(\Gamma)\otimes L^{2}(\mathcal{S}_{\mathscr{L}})\to\mathbb{C}\delta_{\gamma}\otimes L^{2}(\mathcal{S}_{\mathscr{L}})

the canonical projection. Let AA be as given, and write A0(Γ,σ)A\oplus 0\in\mathscr{B}(\Gamma,\sigma) as a matrix (Aij)(A_{ij}), where

Aij=γΓAij,γγ¯,A_{ij}=\sum_{\gamma\in\Gamma}A_{ij,\gamma}\bar{\gamma}, (4.9)

as in (4.6). Letting AγA_{\gamma} be the element of 𝒦{\mathcal{K}} given by the matrix (Aij,γ)i,j(A_{ij,\gamma})_{i,j\in\mathbb{N}}, one observes that

qγ(A0)qe=γ¯Aγ.q_{\gamma}(A\oplus 0)q_{e}=\bar{\gamma}\otimes A_{\gamma}. (4.10)

Now by (4.7), together with the fact that τσ(g)Tr\tau^{(g)}_{\sigma}\otimes\operatorname{Tr} converges absolutely on elements of (Γ,σ)\mathscr{B}(\Gamma,\sigma), we have

(τσ(g)Tr)(A0)\displaystyle(\tau^{(g)}_{\sigma}\otimes\operatorname{Tr})(A\oplus 0) =ih(g)θ(h)Aii,h\displaystyle=\sum_{i\in\mathbb{N}}\sum_{h\in(g)}\theta(h)A_{ii,h}
=h(g)θ(h)Tr(Ah),\displaystyle=\sum_{h\in(g)}\theta(h)\operatorname{Tr}(A_{h}),

Thus by Definition 3.15 it suffices to show that for each h(g)h\in(g),

Tr(Ah)=Tr(c1/2Th1Ac1/2).\operatorname{Tr}(A_{h})=\operatorname{Tr}(c^{1/2}T_{h^{-1}}Ac^{1/2}). (4.11)

To this end, let B={ui}iB=\{u_{i}\}_{i\in\mathbb{N}} be the orthonormal basis of L2(𝒮)L^{2}(\mathcal{S}_{\mathscr{L}}) from the start of this section. By definition,

A0=jσAjσ,A\oplus 0=j_{\sigma}Aj_{\sigma}^{*},

where jσj_{\sigma} is as in (3.3). Hence by (4.10), we have

(h¯Ah)(δeui)\displaystyle(\bar{h}\otimes A_{h})(\delta_{e}\otimes u_{i}) =qhjσAjσ(δeui)\displaystyle=q_{h}\circ j_{\sigma}\circ A\circ j_{\sigma}^{*}(\delta_{e}\otimes u_{i}) (4.12)

for each ii\in\mathbb{N} and hΓh\in\Gamma. The map jσj_{\sigma}^{*} can be written explicitly as follows: given vl2(Γ)L2(𝒮)v\in l^{2}(\Gamma)\otimes L^{2}(\mathcal{S}_{\mathscr{L}}) and xMx\in M,

(jσv)(x)=γΓc(γ1x)Sγ1(γ1x)γ(v(γ,γ1x)).(j_{\sigma}^{*}v)(x)=\sum_{\gamma\in\Gamma}c(\gamma^{-1}x)S_{\gamma^{-1}}(\gamma^{-1}x)\cdot\gamma\cdot(v(\gamma,\gamma^{-1}x)).

Applying this to δeui\delta_{e}\otimes u_{i} and using that Se1S_{e}\equiv 1 gives

jσ(δeui)(x)\displaystyle j_{\sigma}^{*}(\delta_{e}\otimes u_{i})(x) =γΓc1/2(γ1x)Sγ1(γ1x)γ((δe(γ)ui(γ1x)))\displaystyle=\sum_{\gamma\in\Gamma}c^{1/2}(\gamma^{-1}x)S_{\gamma^{-1}}(\gamma^{-1}x)\cdot\gamma\cdot((\delta_{e}(\gamma)u_{i}(\gamma^{-1}x)))
=c1/2(x)ui(x).\displaystyle=c^{1/2}(x)u_{i}(x).

Thus (4.12) equals qhjσA(cui)q_{h}\circ j_{\sigma}\circ A(cu_{i}). By (3.3), jσ(A(cui))(γ)=c1/2Tγ1(A(c1/2ui))j_{\sigma}(A(cu_{i}))(\gamma)=c^{1/2}T_{\gamma^{-1}}(A(c^{1/2}u_{i})), hence

Aγ(ui)=c1/2Tγ1A(c1/2ui).\displaystyle A_{\gamma}(u_{i})=c^{1/2}T_{\gamma^{-1}}A(c^{1/2}u_{i}). (4.13)

It follows that for any h(g)h\in(g),

Tr(Ah)\displaystyle\operatorname{Tr}(A_{h}) =ic1/2Th1A(c1/2ui),uiL2(𝒮)\displaystyle=\sum_{i}\langle c^{1/2}T_{h^{-1}}A(c^{1/2}u_{i}),u_{i}\rangle_{L^{2}(\mathcal{S}_{\mathscr{L}})}
=Tr(c1/2Th1Ac1/2),\displaystyle=\operatorname{Tr}(c^{1/2}T_{h^{-1}}Ac^{1/2}),

which establishes (4.11). ∎

Theorem 4.5.

Let MM, Γ\Gamma, and DD be as in the statement of Theorem 1.5 and Definition 2.4. Let gΓg\in\Gamma be a σ\sigma-regular element for a multiplier σ\sigma. Then

  1. (i)

    etD20(Γ,σ)e^{-tD^{2}}\oplus 0\in\mathscr{B}(\Gamma,\sigma), where (Γ,σ)\mathscr{B}(\Gamma,\sigma) is as in (4.3);

  2. (ii)

    we have

    (τσ(g))IndΓ,σ(D)=strσ(g)(etD2).\big{(}\tau^{(g)}_{\sigma}\big{)}_{*}\operatorname{Ind}_{\Gamma,\sigma}(D)=\textnormal{str}_{\sigma}^{(g)}(e^{-tD^{2}}).
Proof.

For (i), note that by (4.10) and (4.13), we can write

etD20=γΓγ¯c1/2Tγ1etD2c1/2.e^{-tD^{2}}\oplus 0=\sum_{\gamma\in\Gamma}\bar{\gamma}\otimes c^{1/2}T_{\gamma^{-1}}e^{-tD^{2}}c^{1/2}.

Under the identification of L2(𝒮)L^{2}(\mathcal{S}_{\mathscr{L}}) with l2()l^{2}(\mathbb{N}) given at the start of this section, the operator c1/2Tγ1etD2c1/2c^{1/2}T_{\gamma^{-1}}e^{-tD^{2}}c^{1/2} is an element of \mathscr{R}. Hence by Lemma 4.1, it suffices to prove that

γ(γ)kc1/2Tγ1etD2c1/2Δ<\sum_{\gamma}\ell(\gamma)^{k}\|c^{1/2}T_{\gamma^{-1}}e^{-tD^{2}}c^{1/2}\circ\Delta\|<\infty (4.14)

for any k0k\geq 0, where the norm is the operator norm on l2()L2(𝒮)l^{2}(\mathbb{N})\cong L^{2}(\mathcal{S}_{\mathscr{L}}).

Note that by construction, the operator Δ\Delta is local in the sense that for any uL2(𝒮)u\in L^{2}(\mathcal{S}_{\mathscr{L}}), we have supp(Δu)supp(u)\operatorname{supp}(\Delta u)\subseteq\operatorname{supp}(u), while if supp(u)supp(c)\operatorname{supp}(u)\subseteq\operatorname{supp}(c), then Δ\Delta acts on uu as D2D^{2}. Since the operator c1/2Tγ1etD2c1/2c^{1/2}T_{\gamma^{-1}}e^{-tD^{2}}c^{1/2} is supported only on the subset supp(c)×supp(c)M×M\operatorname{supp}(c)\times\operatorname{supp}(c)\subseteq M\times M, and c1/2D2=D2c1/2+DA1+A2c^{1/2}D^{2}=D^{2}c^{1/2}+DA_{1}+A_{2} for some operators A1,A2A_{1},A_{2} of order zero, it follows from the Gaussian decay of the heat kernel that there exist constants C1C_{1}, C2C_{2}, and C3C_{3} such that for all γ\gamma with (γ)>C\ell(\gamma)>C, we have

c1/2Tγ1etD2c1/2D2C2eC3(γ)2.\|c^{1/2}T_{\gamma^{-1}}e^{-tD^{2}}c^{1/2}D^{2}\|\leq C_{2}e^{-C_{3}\ell(\gamma)^{2}}.

Since Γ\Gamma is finitely generated, there exist constants C4,C5C_{4},C_{5} such that the number of group elements γ\gamma with (γ)l\ell(\gamma)\leq l is at most C4eC5lC_{4}e^{C_{5}l}. Setting

C6=(γ)<C1(γ)kc1/2Tγ1etD2c1/2D2C_{6}=\sum_{\ell(\gamma)<C_{1}}\ell(\gamma)^{k}\|c^{1/2}T_{\gamma^{-1}}e^{-tD^{2}}c^{1/2}D^{2}\|

and combining the above observations shows that

γ(γ)kc1/2Tγ1etD2c1/2D2\displaystyle\sum_{\gamma}\ell(\gamma)^{k}\|c^{1/2}T_{\gamma^{-1}}e^{-tD^{2}}c^{1/2}D^{2}\| =C6+(γ)C1(γ)kc1/2Tγ1etD2c1/2D2\displaystyle=C_{6}+\sum_{\ell(\gamma)\geq C_{1}}\ell(\gamma)^{k}\|c^{1/2}T_{\gamma^{-1}}e^{-tD^{2}}c^{1/2}D^{2}\|
C6+(γ)C1(γ)kC4eC5(γ)C2eC3(γ)2,\displaystyle\leq C_{6}+\sum_{\ell(\gamma)\geq C_{1}}\ell(\gamma)^{k}C_{4}e^{C_{5}\ell(\gamma)}C_{2}e^{-C_{3}\ell(\gamma)^{2}},

which is finite. This establishes (4.14).

For (ii), note that we have a commutative diagram

K0(Cr(Γ,σ)𝒦(L2(𝒮))){K_{0}(C^{*}_{r}(\Gamma,\sigma)\otimes\mathcal{K}(L^{2}(\mathcal{S}_{\mathscr{L}})))}K0(Cr(Γ,σ)){K_{0}(C^{*}_{r}(\Gamma,\sigma))}{\mathbb{C}},{\mathbb{C},}Tr\scriptstyle{\operatorname{Tr}_{*}}(τσ(g)Tr)\scriptstyle{(\tau^{(g)}_{\sigma}\otimes\operatorname{Tr})_{*}}(τσ(g))\scriptstyle{(\tau^{(g)}_{\sigma})_{*}}=\scriptstyle{=}            

where the map Tr\operatorname{Tr}_{*} is induced by the operator trace on L2(𝒮)L^{2}(\mathcal{S}_{\mathscr{L}}). It follows from [13, Exercise 12.7.3] that the element

( 0)IndΓ,σ,L2(D)K0(Cr(Γ,σ)𝒦(L2(𝒮)))(\oplus\,0)_{*}\operatorname{Ind}_{\Gamma,\sigma,L^{2}}(D)\in K_{0}(C^{*}_{r}(\Gamma,\sigma)\otimes\mathcal{K}(L^{2}(\mathcal{S}_{\mathscr{L}})))

can be represented explicitly by the following difference of idempotents constructed from the heat operator:

(etDD+0et2DD+1etDD+DD+D0et2D+D1etD+DD+DD+0(1etD+D)0)(0001).\begin{pmatrix}e^{-tD^{-}D^{+}}\oplus 0&e^{-\frac{t}{2}D^{-}D^{+}}\frac{1-e^{-tD^{-}D^{+}}}{D^{-}D^{+}}D^{-}\oplus 0\\ e^{-\frac{t}{2}D^{+}D^{-}}\frac{1-e^{-tD^{+}D^{-}}}{D^{+}D^{-}}D^{+}\oplus 0&(1-e^{-tD^{+}D^{-}})\oplus 0\end{pmatrix}-\begin{pmatrix}0&0\\ 0&1\end{pmatrix}. (4.15)

By part (i), the diagonal entries are in the unitisation of (Γ,σ)\mathscr{B}(\Gamma,\sigma), and a similar argument shows that this is also true of the off-diagonal entries. Thus

(τσ(g))(IndΓ,σ(D))\displaystyle\qquad\quad\big{(}\tau^{(g)}_{\sigma}\big{)}_{*}(\operatorname{Ind}_{\Gamma,\sigma}(D)) =(τσ(g)Tr)( 0)IndΓ,σ,L2(D)\displaystyle=(\tau^{(g)}_{\sigma}\otimes\operatorname{Tr})_{*}(\oplus\,0)_{*}\operatorname{Ind}_{\Gamma,\sigma,L^{2}}(D)
=(τσ(g)Tr)(etDD+0)+(τσ(g)Tr)(etD+D0)\displaystyle=(\tau^{(g)}_{\sigma}\otimes\operatorname{Tr})(e^{-tD^{-}D^{+}}\oplus 0)+(\tau^{(g)}_{\sigma}\otimes\operatorname{Tr})(-e^{-tD^{+}D^{-}}\oplus 0)
=strσ(g)(etD2).\displaystyle=\operatorname{str}^{(g)}_{\sigma}(e^{-tD^{2}}).

where we have used Proposition 4.4 for the last equality.

We now turn our attention to Theorem 1.5. To begin, we have:

Proposition 4.6.

Let MM, Γ\Gamma, DD, and σ\sigma be as in Theorem 1.5. Let ω\omega be as in (2.1). If the Γ\Gamma-invariant Riemannian metric on MM satisfies

infxM(κ(x)4c(ω)x)>0,\inf_{x\in M}(\kappa(x)-4\|c(\omega)\|_{x})>0,

where the norm is taken fibrewise in End(𝒮)\operatorname{End}(\mathcal{S}_{\mathscr{L}}), then

IndΓ,σ,L2(D)=0K0(C(M;L2(𝒮))Γ,σ).\operatorname{Ind}_{\Gamma,\sigma,L^{2}}(D)=0\in K_{0}\big{(}C^{*}(M;L^{2}(\mathcal{S}_{\mathscr{L}}))^{\Gamma,\sigma}\big{)}.

In particular IndΓ,σ(D)=0K0(Cr(Γ,σ)).\operatorname{Ind}_{\Gamma,\sigma}(D)=0\in K_{0}(C^{*}_{r}(\Gamma,\sigma)).

The proof of this proposition uses the Bochner-Lichnerowicz formula.

Lemma 4.7 (Bochner-Lichnerowicz).

Let NN be a spin Riemannian manifold, and let 𝒮NN\mathcal{S}_{N}\to N be the spinor bundle with connection 𝒮N\nabla^{\mathcal{S}_{N}}. Let E\nabla^{E} be a Hermitian connection on a Hermitian vector bundle EME\to M. Then

DE2=+κ4+c(RE),D_{E}^{2}=\nabla^{*}\nabla+\frac{\kappa}{4}+c(R^{E}),

where DED_{E} is the Dirac operator associated to the connection =𝒮N1+1E\nabla=\nabla^{\mathcal{S}_{N}}\otimes 1+1\otimes\nabla^{E} on 𝒮NE\mathcal{S}_{N}\otimes E, κ\kappa is the scalar curvature on NN, and c(RE)c(R^{E}) denotes Clifford multiplication by the curvature of E\nabla^{E}.

Proof of Proposition 4.6.

Let L\nabla^{L} be the Hermitian connection on the trivial bundle M\mathscr{L}\to M defined by iηi\eta, where η\eta is any one-form satisfying dη=ωd\eta=\omega. Applying Lemma 4.7 with DE=∂̸D_{E}=\not{\partial}\otimes\nabla^{\mathscr{L}} gives

(∂̸)2=+κ4+ic(ω),(\not{\partial}\otimes\nabla^{\mathscr{L}})^{2}=\nabla^{*}\nabla+\frac{\kappa}{4}+ic(\omega),

where =𝒮1+1\nabla=\nabla^{\mathcal{S}}\otimes 1+1\otimes\nabla^{\mathscr{L}}. Thus for all uL2(M,𝒮)u\in L^{2}(M,\mathcal{S}_{\mathscr{L}}), we have

D2u,uL2\displaystyle\langle D^{2}u,u\rangle_{L^{2}} =u,uL2+14κu,uL2+ic(ω)u,uL20\displaystyle=\langle\nabla u,\nabla u\rangle_{L^{2}}+\frac{1}{4}\langle\kappa u,u\rangle_{L^{2}}+\langle ic(\omega)u,u\rangle_{L^{2}}\geq 0

by our assumption. Then IndΓ,σ,L2(D)\operatorname{Ind}_{\Gamma,\sigma,L^{2}}(D) can be defined by choosing the normalising function χ\chi equal to the sign function, and a routine computation then shows that the index representative (6.4) is exactly zero. The final statement follows by applying the homomorphism ( 0)(\oplus\,0)_{*} to IndΓ,σ,L2(D)\operatorname{Ind}_{\Gamma,\sigma,L^{2}}(D), as in (3.22). ∎

Corollary 4.8.

Let MM and Γ\Gamma be as in Theorem 1.5. For each ss\in\mathbb{R}, let DsD^{s} be the twisted Dirac operator constructed in subsection 2.2. If the metric on MM has positive scalar curvature, then for ss sufficiently small we have IndΓ,σs,L2(Ds)=0\operatorname{Ind}_{\Gamma,\sigma^{s},L^{2}}(D^{s})=0.

Proof.

The curvature of the connection ,s\nabla^{\mathscr{L},s} is isωis\omega. Thus by Proposition 4.6, it suffices to show that

infxM(κ(x)4sc(ω)x)>0\inf_{x\in M}(\kappa(x)-4\|sc(\omega)\|_{x})>0 (4.16)

for ss sufficiently small. For xMx\in M, let iλ1(x),,iλn(x)i\lambda_{1}(x),\ldots,i\lambda_{n}(x) be the pointwise eigenvalues of ω\omega, which we view as a skew-symmetric endomorphism of TMTM using the Riemannian metric. Since the action of Γ\Gamma on MM is cocompact, κ(x)\kappa(x) is uniformly bounded below by some κ0>0\kappa_{0}>0, while there exists a constant CC such that

c(ω)xCj=1n|λj(x)|\|c(\omega)\|_{x}\leq C\cdot\sum_{j=1}^{n}|\lambda_{j}(x)|

for each xx. Since ω\omega is Γ\Gamma-invariant, the right-hand side is again uniformly bounded over MM. It follows that (4.16) holds for ss sufficiently small. ∎

Proposition 4.9.

Let (g)Γ(g)\subseteq\Gamma be a σs\sigma^{s}-regular conjugacy class, with σs\sigma^{s} as in (2.6). Then the operators etDsD+se^{-tD^{s}_{-}D^{s}_{+}} and etD+sDse^{-tD^{s}_{+}D^{s}_{-}} are of (σs,g)(\sigma^{s},g)-trace class.

Proof.

Fix ss\in\mathbb{R} and ϕ1,ϕ2Cc(M)\phi_{1},\phi_{2}\in C_{c}(M). Let Ks,tK_{s,t} be the Schwartz kernel of et(Ds)2e^{-t(D^{s})^{2}}. By standard estimates for the heat kernel on cocompact manifolds (see for example [24, Corollary 3.5]), there exists a constant C2C_{2}, depending on tt, such that

h(g)h1Ks,t(hx,x)x<C2\sum_{h\in(g)}\|h^{-1}K_{s,t}(hx,x)\|_{x}<C_{2}

for all xMx\in M, where the norm is taken in End(𝒮)\operatorname{End}(\mathcal{S}_{\mathscr{L}}). Then

h(g)|θ(h)Str(ϕ1Th1et(Ds)2ϕ2)|\displaystyle\sum_{h\in(g)}|\theta(h)\cdot\operatorname{Str}(\phi_{1}T_{h^{-1}}e^{-t(D^{s})^{2}}\phi_{2})| h(g)|Str(ϕ1Th1et(Ds)2ϕ2)|\displaystyle\leq\sum_{h\in(g)}|\operatorname{Str}(\phi_{1}T_{h^{-1}}e^{-t(D^{s})^{2}}\phi_{2})|
C1h(g)Mϕ1(x)ϕ2(x)h1Ks,t(hx,x)x𝑑x\displaystyle\leq C_{1}\sum_{h\in(g)}\int_{M}\phi_{1}(x)\phi_{2}(x)\|h^{-1}K_{s,t}(hx,x)\|_{x}\,dx
C1C2ϕ1ϕ2L1.\displaystyle\leq C_{1}C_{2}\|\phi_{1}\phi_{2}\|_{L^{1}}.

for some constant C1C_{1}, by Fubini’s theorem. Hence et(Ds)2e^{-t(D^{s})^{2}} is of (σs,g)(\sigma^{s},g)-trace class. ∎

Next, let ΩΓ\Omega\subseteq\Gamma be a subset such that

  1. (i)

    for all h(g)h\in(g), there exists some kΩk\in\Omega such that k1gk=hk^{-1}gk=h;

  2. (ii)

    if k1,k2Ωk_{1},k_{2}\in\Omega and k1k2k_{1}\neq k_{2}, then k11gk1k21gk2k_{1}^{-1}gk_{1}\neq k_{2}^{-1}gk_{2}.

Let cc be a cut-off function for the action of Γ\Gamma on MM. One checks that the function cgCc(M)c^{g}\in C_{c}^{\infty}(M) defined by

cg(x)=kΩc(k1x)c^{g}(x)=\sum_{k\in\Omega}c(k^{-1}x) (4.17)

is a cut-off function for the action of ZgZ^{g} on MM.

Proposition 4.10.

Let α\alpha and σ\sigma be as in (2.5) and (2.6). Let (g)Γ(g)\subseteq\Gamma be the conjugacy class of an α\alpha-regular element and cgc^{g} as in (4.17). Then

strσs(g)et(Ds)2=Meisϕg(x)cg(x)Str(g1Ks,t(gx,x))𝑑x,\operatorname{str}^{(g)}_{\sigma^{s}}e^{-t(D^{s})^{2}}=\int_{M}e^{-is\phi_{g}(x)}c^{g}(x)\operatorname{Str}(g^{-1}K_{s,t}(gx,x))\,dx,

where Ks,tK_{s,t} is the Schwartz kernel of et(Ds)2e^{-t(D^{s})^{2}}.

Proof.

We begin by establishing a useful algebraic identity:

θs(k1gk)eis(ϕk1gk(k1x)+ϕk(k1gx)ϕk(k1x))=eisϕg(x),\theta^{s}(k^{-1}gk)e^{-is(\phi_{k^{-1}gk}(k^{-1}x)+\phi_{k}(k^{-1}gx)-\phi_{k}(k^{-1}x))}=e^{-is\phi_{g}(x)}, (4.18)

where θs\theta^{s} is as in (3.6).

It suffices to establish this identity for s=1s=1. The case of general ss follows by replacing θ\theta and ϕ\phi by θs\theta^{s} and sϕs\phi respectively, similar to the proof of Corollary 3.11. To proceed, note that by (2.3) applied with xk1xx\to k^{-1}x, γk1gk\gamma\to k^{-1}gk, and γk\gamma^{\prime}\to k, the function

xϕk1gk(k1x)ϕk(k1gx)+ϕgk(k1x)x\mapsto-\phi_{k^{-1}gk}(k^{-1}x)-\phi_{k}(k^{-1}gx)+\phi_{gk}(k^{-1}x)

is constant on MM, hence

ei(ϕk1gk(k1x)+ϕk(k1gx))=eiϕgk(k1x)C1e^{-i(\phi_{k^{-1}gk}(k^{-1}x)+\phi_{k}(k^{-1}gx))}=e^{-i\phi_{gk}(k^{-1}x)}\cdot C_{1} (4.19)

for some C1U(1)C_{1}\in\operatorname{U}(1). Letting x=kx0x=kx_{0} and using that ϕe0\phi_{e}\equiv 0 gives

C1=eiϕk(k1gkx0)=σ¯(k,k1gk).C_{1}=e^{-i\phi_{k}(k^{-1}gkx_{0})}=\bar{\sigma}(k,k^{-1}gk).

Now by the cocycle identity and Lemma 3.4,

θ(k1gk)\displaystyle\theta(k^{-1}gk) =σ(g1,k)σ(k1,g1k)\displaystyle=\sigma(g^{-1},k)\sigma(k^{-1},g^{-1}k)
=σ(k1,g1)σ(k1g1k)\displaystyle=\sigma(k^{-1},g^{-1})\sigma(k^{-1}g^{-1}k)
=σ¯(g,k)σ(k1g1,k).\displaystyle=\bar{\sigma}(g,k)\sigma(k^{-1}g^{-1},k).

Using this and (4.21), the claimed equality (4.18) becomes

σ¯(g,k)σ(k1g1,k)σ¯(k,k1gk)ei(ϕgk(k1x)+ϕk(k1x))=eiϕg(x).\bar{\sigma}(g,k)\sigma(k^{-1}g^{-1},k)\bar{\sigma}(k,k^{-1}gk)e^{-i(\phi_{gk}(k^{-1}x)+\phi_{k}(k^{-1}x))}=e^{-i\phi_{g}(x)}. (4.20)

Applying (2.3) with xk1xx\to k^{-1}x, γk\gamma\to k, and γg\gamma^{\prime}\to g now shows that the function

xϕk(k1x)ϕg(x)+ϕgk(k1x)x\mapsto-\phi_{k}(k^{-1}x)-\phi_{g}(x)+\phi_{gk}(k^{-1}x)

is constant on MM, hence

ei(ϕgk(k1x)+ϕk(k1x))=eiϕg(x)C2e^{-i(\phi_{gk}(k^{-1}x)+\phi_{k}(k^{-1}x))}=e^{-i\phi_{g}(x)}\cdot C_{2} (4.21)

for some C2U(1)C_{2}\in\operatorname{U}(1). Letting x=kx0x=kx_{0} shows that

C2=eiϕg(kx0)=σ(g,k).C_{2}=e^{i\phi_{g}(kx_{0})}=\sigma(g,k).

This, together with the computation

σ(k1g1,k)σ¯(k,k1gk)=σ(k1g1,k)σ(k1g1k,k1)=1,\displaystyle\sigma(k^{-1}g^{-1},k)\bar{\sigma}(k,k^{-1}gk)=\sigma(k^{-1}g^{-1},k)\sigma(k^{-1}g^{-1}k,k^{-1})=1,

establishes (4.20) and therefore (4.18).

Next, by Lemma 3.17 and a computation similar to (3.1), we can use the set Ω\Omega defined above to write strσs(g)et(Ds)2\operatorname{str}^{(g)}_{\sigma^{s}}e^{-t(D^{s})^{2}} as

kΩθs(k1gk)Meis(ϕk1gk(k1x)+ϕk(k1gx)ϕk(k1x))c(k1x)Str(g1Ks,t(gx,x))𝑑x.\displaystyle\sum_{k\in\Omega}\theta^{s}(k^{-1}gk)\int_{M}e^{-is(\phi_{k^{-1}gk}(k^{-1}x)+\phi_{k}(k^{-1}gx)-\phi_{k}(k^{-1}x))}c(k^{-1}x)\operatorname{Str}(g^{-1}K_{s,t}(gx,x))\,dx.

Similarly to the proof of Proposition 4.9, we can take the sum inside the integral. Then by (4.18) and the definition of cgc^{g} given by (4.17), this is equal to

Meisϕg(x)cg(x)Str(g1Ks,t(gx,x))𝑑x.\displaystyle\int_{M}e^{-is\phi_{g}(x)}c^{g}(x)\operatorname{Str}(g^{-1}K_{s,t}(gx,x))\,dx.

We can now finish the proof of Theorem 1.5.

Proof of Theorem 1.5.

For (i), first note that

(Ds)2et(Ds)2=12[Ds,Dset(Ds)2]s.(D^{s})^{2}e^{-t(D^{s})^{2}}=\frac{1}{2}[D^{s},D^{s}e^{-t(D^{s})^{2}}]_{s}.

It follows from Corollary 3.24 that

d(strσs(g)et(Ds)2)dt\displaystyle\frac{d(\operatorname{str}^{(g)}_{\sigma^{s}}e^{-t(D^{s})^{2}})}{dt} =strσs(g)((Ds)2et(Ds)2)=0,\displaystyle=-\operatorname{str}_{\sigma^{s}}^{(g)}\big{(}(D^{s})^{2}e^{-t(D^{s})^{2}}\big{)}=0, (4.22)

hence the function tstrσs(g)et(Ds)2t\mapsto\operatorname{str}^{(g)}_{\sigma^{s}}e^{-t(D^{s})^{2}} is constant in t>0t>0. Let Ks,tK_{s,t} be the Schwartz kernel of the operator et(Ds)2e^{-t(D^{s})^{2}}. By standard heat kernel estimates, in the limit t0t\to 0 the integral

Meisϕg(x)cg(x)Str(g1Ks,t(gx,x))𝑑x\int_{M}e^{-is\phi_{g}(x)}c^{g}(x)\operatorname{Str}(g^{-1}K_{s,t}(gx,x))\,dx

localises to arbitrarily small neighbourhoods of the fixed-point submanifold MgM^{g}; see for example [14, Lemma 4.10]. Pick a sufficiently small tubular neighbourhood and identify it with the normal bundle 𝒩\mathcal{N} of MgM^{g} in MM. Since the curvature of the twisting bundle M\mathscr{L}\to M is Γ\Gamma-invariant, the standard asymptotic expansion of Ks,tK_{s,t} (see [3, Theorem 6.11]) applies. The same argument as in the compact case (see [3, Theorem 6.16]) then shows that the above integral equals

MgeisϕgcgA^(Mg)esω/2πidet(1geR𝒩/2πi)1/2,\int_{M^{g}}e^{-is\phi_{g}}c^{g}\cdot\frac{\widehat{A}(M^{g})\cdot e^{-s\omega/2\pi i}}{\det(1-ge^{-R^{\mathcal{N}}/2\pi i})^{1/2}}, (4.23)

where we have used that cg|Mgc^{g}|_{M^{g}} is a cut-off function for the action of ZgZ^{g} on MgM^{g}.

By Remark 2.2, eisϕg(x)e^{-is\phi_{g}}(x) is constant on each connected component of MgM^{g}. Further, the support of cgc^{g} is compact and thus intersects only finitely many of these connected components, M1g,,MNgM^{g}_{1},\ldots,M^{g}_{N}. For each j=1,,Nj=1,\ldots,N, pick a point xjMjx_{j}\in M_{j}, and let 𝒩j\mathcal{N}_{j} be the restriction of 𝒩\mathcal{N} to MjgM^{g}_{j}. By Theorem 4.5 (ii), together with the above discussion, we have

(τσs(g))IndΓ,σs(Ds)\displaystyle(\tau^{(g)}_{\sigma^{s}})_{*}\operatorname{Ind}_{\Gamma,\sigma^{s}}(D^{s}) =strσs(g)(et(Ds)2)\displaystyle=\operatorname{str}^{(g)}_{\sigma^{s}}(e^{-t(D^{s})^{2}})
=j=1mMjgeisϕg(xj)cgA^(Mjg)esω/2πi|Mjgdet(1geR𝒩j/2πi)1/2\displaystyle=\sum_{j=1}^{m}\int_{M^{g}_{j}}e^{-is\phi_{g}(x_{j})}c^{g}\cdot\frac{\widehat{A}(M^{g}_{j})\cdot e^{-s\omega/2\pi i}|_{M^{g}_{j}}}{\det(1-ge^{-R^{\mathcal{N}_{j}}/2\pi i})^{1/2}}
=k=0dimM/2(s)kk!j=1mMjgcgA^(Mg)(iϕg(xj)+ω2πi)k|Mjgdet(1geR𝒩j/2πi)1/2.\displaystyle=\sum_{k=0}^{\dim M/2}\frac{(-s)^{k}}{k!}\sum_{j=1}^{m}\int_{M^{g}_{j}}c^{g}\cdot\frac{\widehat{A}(M^{g})\cdot\left(i\phi_{g}(x_{j})+\frac{\omega}{2\pi i}\right)^{k}|_{M^{g}_{j}}}{\det(1-ge^{-R^{\mathcal{N}_{j}}/2\pi i})^{1/2}}.

For (ii), observe that we can, by a straightforward suspension argument, reduce to the case where MM is even-dimensional. If MM admits a Γ\Gamma-invariant metric of positive scalar curvature, then IndΓ,σsDs\operatorname{Ind}_{\Gamma,\sigma^{s}}D^{s} vanishes for all s(0,δ)s\in(0,\delta) for some 0<δ0<\delta, by Proposition 4.8. Since the map

s(τσs(g))(IndΓ,σsDs)s\mapsto(\tau^{(g)}_{\sigma^{s}})_{*}(\operatorname{Ind}_{\Gamma,\sigma^{s}}D^{s})

is a polynomial in ss, it must vanish identically on \mathbb{R}, hence

j=1mMjgcgA^(Mg)(ϕg(xj)ω2π)k|Mjgdet(1geR𝒩j/2πi)1/2=0\sum_{j=1}^{m}\int_{M^{g}_{j}}c^{g}\cdot\frac{\widehat{A}(M^{g})\cdot\left(\phi_{g}(x_{j})-\frac{\omega}{2\pi}\right)^{k}|_{M^{g}_{j}}}{\det(1-ge^{-R^{\mathcal{N}_{j}}/2\pi i})^{1/2}}=0 (4.24)

for each k0k\geq 0, where we may also assume that each MjgM^{g}_{j} is even-dimensional.

In the case that MgM^{g} is connected, we may choose the point x0x_{0} from (2.4) to lie in MgM^{g}, whence Remark 2.2 implies that eisϕg1e^{-is\phi_{g}}\equiv 1 on MgM^{g}. The formula (4.24) now simplifies to

A^g(M,ωk)=MgcgA^(Mg)ωk|Mgdet(1geR𝒩/2πi)1/2=0\widehat{A}_{g}(M,\omega^{k})=\int_{M^{g}}c^{g}\cdot\frac{\widehat{A}(M^{g})\cdot\omega^{k}|_{M^{g}}}{\det(1-ge^{-R^{\mathcal{N}}/2\pi i})^{1/2}}=0

for each k0k\geq 0. ∎

Remark 4.11.

It should be possible to derive part (ii) of Theorem 1.5 independently of any growth assumptions on (g)(g) by using methods similar to the proof of [19, Theorem 3.4]; see also [27, Remark A.2]. In particular, this would imply that if the Baum-Connes conjecture holds for Γ\Gamma, then the map (τσ(g))(\tau^{(g)}_{\sigma})_{*} from (4.8) can be defined via the σ\sigma-weighted (g)(g)-supertrace strσ(g)\operatorname{str}^{(g)}_{\sigma} from (3.16).

Proof of Corollary 1.6.

By a suspension argument, we may assume without loss of generality that both MM and MgM^{g} are even-dimensional. Further, since the vanishing property is preserved under sums of forms, we may to restrict our attention to the case where

ω=ω1ω2ωm,\omega=\omega_{1}\omega_{2}\ldots\omega_{m},

for some mdimM/2m\leq\dim M/2, where each ωi\omega_{i} is the lift of a differential form on M/ΓM/\Gamma representing a class in fH2(B¯Γ,)f^{*}H^{2}(\underline{B}\Gamma,\mathbb{R}). For each i=1,,mi=1,\ldots,m, let sis_{i}\in\mathbb{R}. Then applying to argument from the proof of Theorem 1.5 to i=1msiωi\sum_{i=1}^{m}s_{i}\omega_{i} instead of sωs\omega, with the twisted Dirac operator defined accordingly, shows that if MM admits a Γ\Gamma-invariant metric of positive scalar curvature, then there exists a δ>0\delta>0 such that

k=0dimM/21k!MgcgA^(Mg)(s1ω1++smωm2πi)k|Mgdet(1geR𝒩/2πi)1/2=0\sum_{k=0}^{\dim M/2}\frac{1}{k!}\int_{M^{g}}c^{g}\cdot\frac{\widehat{A}(M^{g})\cdot\left(\frac{s_{1}\omega_{1}+\cdots+s_{m}\omega_{m}}{2\pi i}\right)^{k}|_{M^{g}}}{\det(1-ge^{-R^{\mathcal{N}}/2\pi i})^{1/2}}=0

whenever si(0,δ)s_{i}\in(0,\delta) for all ii. Since the left-hand side is a polynomial in the variables s1,,sms_{1},\ldots,s_{m}, it vanishes identically on m\mathbb{R}^{m}. In particular, the coefficient of s1s2sms_{1}s_{2}\ldots s_{m} is zero, and this is equal to

(2πi)mm!MgcgA^(Mg)ω|Mgdet(1geR𝒩/2πi)1/2.\frac{(2\pi i)^{-m}}{m!}\int_{M^{g}}c^{g}\cdot\frac{\widehat{A}(M^{g})\cdot\omega|_{M^{g}}}{\det(1-ge^{-R^{\mathcal{N}}/2\pi i})^{1/2}}.

This concludes the proof. ∎

4.1. Special cases

We discuss two special cases of Theorem 1.5 that have already appeared elsewhere in the literature. These correspond to the extreme cases when either the conjugacy class or the multiplier is trivial.

4.1.1. Free action and (g)=(e)(g)=(e)

In this case, the canonical trace

τσ(e):σΓ\displaystyle\tau_{\sigma}^{(e)}\colon\mathbb{C}^{\sigma}\Gamma \displaystyle\to\mathbb{C}
γΓaγγ¯\displaystyle\sum_{\gamma\in\Gamma}a_{\gamma}\bar{\gamma} ae\displaystyle\mapsto a_{e} (4.25)

extends continuously to a trace Cr(Γ)C^{*}_{r}(\Gamma)\to\mathbb{C} and induces a linear map

(τσ(e)):K0(Cr(Γ,σ)).\big{(}\tau_{\sigma}^{(e)}\big{)}_{*}\colon K_{0}(C^{*}_{r}(\Gamma,\sigma))\to\mathbb{C}.

Since the action of Γ\Gamma on MM is free, we may work with the classifying space BΓB\Gamma instead of B¯Γ\underline{B}\Gamma, as done in [17]. Let f:M/ΓBΓf\colon M/\Gamma\to B\Gamma be the classifying map for MM, and let [β]H2(BΓ,)[\beta]\in H^{2}(B\Gamma,\mathbb{R}). Let ω0\omega_{0} be a differential form on the quotient manifold M/ΓM/\Gamma such that f[β]=[ω0]f^{*}[\beta]=[\omega_{0}], and let ω\omega be the Γ\Gamma-invariant lift of ω0\omega_{0} to MM.

Since ϕe0\phi_{e}\equiv 0, Theorem 1.5 (i) reduces to the twisted L2L^{2}-index theorem in of Mathai [17, Theorem 3.6] for the case of the spin-Dirac operator, namely that

(τσ(e))IndΓ,σ(D)=McA^(M)eω/2πi=M/ΓA^(M/Γ)eω0/2πi,\big{(}\tau^{(e)}_{\sigma}\big{)}_{*}\operatorname{Ind}_{\Gamma,\sigma}(D)=\int_{M}c\cdot\widehat{A}(M)\cdot e^{-\omega/2\pi i}=\int_{M/\Gamma}\widehat{A}(M/\Gamma)\cdot e^{-\omega_{0}/2\pi i},

where cc is a cut-off function for the Γ\Gamma-action on MM. Theorem 1.5 (ii) recovers the result if the M/ΓM/\Gamma admits a metric of positive scalar curvature, then for each non-negative integer kk,

MA^(M/Γ)ω0k=0.\int_{M}\widehat{A}(M/\Gamma)\cdot\omega_{0}^{k}=0.

This is [17, Corollary 1].

4.1.2. Trivial multiplier σ1\sigma\equiv 1

In this case, the projectively invariant Dirac operator DD is simply the Γ\Gamma-invariant Dirac operator acting on sections of the spinor bundle 𝒮M\mathcal{S}\to M. The element IndΓ,σ(D)\operatorname{Ind}_{\Gamma,\sigma}(D) reduces to the usual Γ\Gamma-equivariant higher index IndΓ(D)K0(Cr(Γ))\operatorname{Ind}_{\Gamma}(D)\in K_{0}(C^{*}_{r}(\Gamma)) for cocompact actions [2]. The trace τσ(g)\tau^{(g)}_{\sigma} is then the unweighted trace τ(g):Γ\tau^{(g)}\colon\mathbb{C}\Gamma\to\mathbb{C} from (3).

Theorem 1.5 (i) then states that if (g)(g) has polynomial growth, then

τ(g)IndΓ(D)=MgcgA^(Mg)det(1geR𝒩/2πi)1/2,\tau^{(g)}_{*}\operatorname{Ind}_{\Gamma}(D)=\int_{M^{g}}c^{g}\cdot\frac{\widehat{A}(M^{g})}{\det(1-ge^{-R^{\mathcal{N}}/2\pi i})^{1/2}},

where we have implicitly taken ω\omega to be zero. This recovers the formula [24, Theorem 6.1] in the case of the spin-Dirac operator.


5. A neighbourhood PSC obstruction

In this section we prove Theorem 1.7. We will make use of a Callias-type index theorem for projectively invariant operators.

5.1. Projectively invariant Callias-type operators

Let us begin with a general definition and discussion of projectively invariant Callias-type operators and their higher indices.

Let MM be a complete Riemannian manifold on which a discrete gorup Γ\Gamma acts properly and isometrically, and suppose that H1(M)=0H^{1}(M)=0. Let

f:M/ΓB¯Γf\colon M/\Gamma\to\underline{B}\Gamma

be the classifying map of MM. Let [β]H2(B¯Γ,)[\beta]\in H^{2}(\underline{B}\Gamma,\mathbb{R}), and let [ω0]=f[β][\omega_{0}]=f^{*}[\beta] in the de Rham cohomology of the orbifold M/ΓM/\Gamma. Let ω\omega be the Γ\Gamma-invariant lift of ω0\omega_{0} to MM. As in subsection 2.2, we obtain a one-form η\eta, a family ϕ\phi of functions on MM, and a family of multipliers σs\sigma^{s} on Γ\Gamma parameterised by ss\in\mathbb{R}.

Remark 5.1.

The cocycles α\alpha and σs\sigma^{s} from (2.5) and (2.6) can be defined equivalently in terms of the above data restricted to any submanifold PMP\subseteq M preserved by the action of Γ\Gamma. Indeed, by working with the de Rham differential on PP instead of MM, (2.3) implies that the family

ϕ|P{ϕγ|P:γΓ}\phi|_{P}\coloneqq\{\phi_{\gamma}|_{P}\colon\gamma\in\Gamma\}

satisfies

dP(ϕγ|P+γ1ϕγ|Pϕγγ|P)=0.d_{P}(\phi_{\gamma}|_{P}+\gamma^{-1}\phi_{\gamma^{\prime}}|_{P}-\phi_{\gamma^{\prime}\gamma}|_{P})=0.

It follows that α\alpha and σs\sigma^{s} can be defined equivalently as

α(γ,γ)=12π(ϕγ|P(x)+ϕγ|P(γx)ϕγγ|P(x)),\alpha(\gamma,\gamma^{\prime})=\frac{1}{2\pi}(\phi_{\gamma}|_{P}(x)+\phi_{\gamma}|_{P}(\gamma^{\prime}x)-\phi_{\gamma^{\prime}\gamma}|_{P}(x)),
σs(γ,γ)=e2πisα(γ,σ).\sigma^{s}(\gamma,\gamma^{\prime})=e^{2\pi is\alpha(\gamma,\sigma)}.

Since the discussion in rest of this subsection applies uniformly to σs\sigma^{s} for any ss with only minor and obvious adjustments, let us now fix s=1s=1 and write σ=σ1\sigma=\sigma^{1}.

Let EME\to M be a 2\mathbb{Z}_{2}-graded Clifford bundle over MM such that L2(E)L^{2}(E) is equipped with the projective action TT from Definition 2.3. Let DD be an odd-graded Dirac operator acting on smooth sections of EE that commutes with TT.

Definition 5.2.

An odd-graded, Γ\Gamma-equivariant fibrewise Hermitian bundle endomorphism Φ\Phi of EE is admissible for DD if DΦ+ΦDD\Phi+\Phi D is an endomorphism of EE such that there exists a cocompact subset ZMZ\subseteq M and a constant C>0C>0 such that the pointwise estimate

Φ2DΦ+ΦD+C\Phi^{2}\geq\|D\Phi+\Phi D\|+C (5.1)

holds over MZM\setminus Z. In this setting, D+ΦD+\Phi is called a (Γ,σ)(\Gamma,\sigma)-invariant Callias-type operator.

Remark 5.3.

A bundle endomorphism on L2(E)L^{2}(E) commutes with the projective action TT from Definition 2.3 if and only if it commutes with the unitary action UU of the group.

A key property of the operator D+ΦD+\Phi is that it has an index in K(Cr(Γ,σ))K_{*}(C^{*}_{r}(\Gamma,\sigma)). The indices of such operators can be defined either via Roe algebras, similar to what was done in subsection 3.3, or using Hilbert Cr(Γ,σ)C^{*}_{r}(\Gamma,\sigma)-modules. We will take the latter approach in order to frame the discussion in parallel with that in [11] for the untwisted case.

To this end, given elements a=γΓaγγσΓa=\sum_{\gamma\in\Gamma}a_{\gamma}\gamma\in\mathbb{C}^{\sigma}\Gamma and sections s,s1,s2Cc(E)s,s_{1},s_{2}\in C_{c}(E), the formulas

(sa)(x)\displaystyle(s\cdot a)(x) γΓaγ(Tγ1s)(x)\displaystyle\coloneqq\sum_{\gamma\in\Gamma}a_{\gamma}(T_{\gamma^{-1}}s)(x)
(s1,s2)(γ)\displaystyle(s_{1},s_{2})(\gamma) (s1,Tγs2)\displaystyle\coloneqq(s_{1},T_{\gamma}s_{2}) (5.2)

define a pre-Hilbert σΓ\mathbb{C}^{\sigma}\Gamma-module structure on Cc(E)C_{c}(E).

Definition 5.4.

Let σ\mathcal{E}^{\sigma} be the Hilbert Cr(Γ,σ)C^{*}_{r}(\Gamma,\sigma)-module completion of Cc(E)C_{c}(E) with respect to (5.1).

The admissibility condition (5.1) implies that the operator D+ΦD+\Phi is projectively Fredholm in the following sense:

Proposition 5.5.

There exists a cocompactly supported, GG-invariant continuous function ff on MM such that

F(D+Φ)((D+Φ)2+f)1/2(σ),F\coloneqq(D+\Phi)\bigl{(}(D+\Phi)^{2}+f\bigr{)}^{-1/2}\in{\mathcal{B}}(\mathcal{E}^{\sigma}), (5.3)

such that the pair (σ,F)(\mathcal{E}^{\sigma},F) is a cycle in KK(,Cr(Γ,σ))K\!K(\mathbb{C},C^{*}_{r}(\Gamma,\sigma)). The class [σ,F][\mathcal{E}^{\sigma},F] is independent of the choice of ff.

Proof.

The proof is analogous to that of [10, Theorem 4.19]. Instead of the Hilbert module C(G)C^{*}(G)-module \mathcal{E} used there, we work with σ\mathcal{E}^{\sigma}. ∎

Definition 5.6.

The (Γ,σ)(\Gamma,\sigma)-index of D+ΦD+\Phi is the class

IndΓ,σ(D+Φ)[σ,F]K0(Cr(Γ,σ))KK(,Cr(Γ,σ)).\operatorname{Ind}_{\Gamma,\sigma}(D+\Phi)\coloneqq[\mathcal{E}^{\sigma},F]\in K_{0}(C^{*}_{r}(\Gamma,\sigma))\cong K\!K(\mathbb{C},C^{*}_{r}(\Gamma,\sigma)).

5.2. Localisation of projective Callias-type indices

One of the key properties of Callias-type operators in the equivariant setting is that the their indices can be calculated by localising to a cocompact subset of the manifold [11, Theorem 3.4]. In the projective setting, a similar result holds. For this, we will assume that the (Γ,σ)(\Gamma,\sigma)-invariant Dirac operator DD from Definition 5.2 takes the form of a Dirac operator twisted by a line bundle, as in Definition 2.4.

Let E0E_{0} be an ungraded Γ\Gamma-equivariant Clifford bundle over MM equipped with a Γ\Gamma-invariant Hermitian connection E0\nabla^{E_{0}}. Define the Hermitian connection

=d+iη\nabla^{\mathscr{L}}=d+i\eta (5.4)

on a Γ\Gamma-equivariantly trivial Hermitian line bundle M\mathscr{L}\to M, and form the connection E0,=E01+1\nabla^{E_{0,\mathscr{L}}}=\nabla^{E_{0}}\otimes 1+1\otimes\nabla^{\mathscr{L}} on the bundle

E0,E0.E_{0,\mathscr{L}}\coloneqq E_{0}\otimes\mathscr{L}. (5.5)

In the notation of subsection 5.1, we will take

E=EE0,E0,E=E_{\mathscr{L}}\coloneqq E_{0,\mathscr{L}}\oplus E_{0,\mathscr{L}}

where the first copy of E0,E_{0,\mathscr{L}} is given the even grading, and the second copy the odd grading. Let D0D_{0} be the Dirac operator on E0,E_{0,\mathscr{L}} associated to E0,\nabla^{E_{0,\mathscr{L}}}, and define

D=(0D0D00)D=\begin{pmatrix}0&D_{0}\\ D_{0}&0\end{pmatrix} (5.6)

on EE_{\mathscr{L}}. Let Φ0\Phi_{0} be Γ\Gamma-invariant a Hermitian endomorphism of E0E_{0} such that Φ2DΦ+ΦD+C\Phi^{2}\geq\|D\Phi+\Phi D\|+C holds outside a cocompact subset ZMZ\subseteq M for some C>0C>0. Let

Φ=(0iΦ01iΦ010).\Phi=\begin{pmatrix}0&i\Phi_{0}\otimes 1\\ -i\Phi_{0}\otimes 1&0\end{pmatrix}. (5.7)

Then D+ΦD+\Phi is a (Γ,σ)(\Gamma,\sigma)-invariant Callias-type operator acting on sections of EE_{\mathscr{L}}, in the sense of Definition 5.2.

Let MMM_{-}\subseteq M be a Γ\Gamma-invariant, cocompact subset containing ZZ in its interior, such that NMN\coloneqq\partial M_{-} is a (not necessary connected) smooth submanifold of MM. Let M+M_{+} be the closure of the complement of MM_{-}, so that N=MM+N=M_{-}\cap M_{+} and M=MM+M=M_{-}\cup M_{+}. We will use the notation

M=MNM+.M=M_{-}\cup_{N}M_{+}.

Let E0,NE^{N}_{0,\mathscr{L}} denote the restriction of E0,E_{0,\mathscr{L}} to NN, equipped with the restricted connection E0,N\nabla^{E_{0,\mathscr{L}}^{N}}. By (5.1), the restriction of Φ0\Phi_{0} to NN is fibrewise invertible. Let

E0,,±NE0,NE^{N}_{0,\mathscr{L},\pm}\subseteq E^{N}_{0,\mathscr{L}}

be the positive and negative eigenbundles of Φ0\Phi_{0}. Clifford multiplication by ii times the unit normal vector field n^\widehat{n} to NN pointing into M+M_{+} defines Γ\Gamma-invariant gradings on both E0,,+NE^{N}_{0,\mathscr{L},+} and E0,,NE^{N}_{0,\mathscr{L},-}. Define the connections

E0,,±Np±E0,Np±\nabla^{E^{N}_{0,\mathscr{L},\pm}}\coloneqq p_{\pm}\nabla^{E_{0,\mathscr{L}}^{N}}p_{\pm} (5.8)

on E0,,±NE^{N}_{0,\mathscr{L},\pm}, where p±:E0,NE0,,±Np_{\pm}\colon E^{N}_{0,\mathscr{L}}\to E^{N}_{0,\mathscr{L},\pm} are the orthogonal projections. Along with the Clifford action of TM|NTM|_{N} on E0,,±NE^{N}_{0,\mathscr{L},\pm}, these connections give rise to two Dirac operators

DE0,,+NandDE0,,N,D^{E^{N}_{0,\mathscr{L},+}}\quad\textnormal{and}\quad D^{E^{N}_{0,\mathscr{L},-}}, (5.9)

both odd-graded, acting on sections of E0,,+NE^{N}_{0,\mathscr{L},+} and E0,,NE^{N}_{0,\mathscr{L},-} respectively.

For each group element γΓ\gamma\in\Gamma, define unitary operators UγN,±,SγN,±U^{N,\pm}_{\gamma},S^{N,\pm}_{\gamma}, and TγN,±T^{N,\pm}_{\gamma} on L2(E0,,±N)L^{2}(E^{N}_{0,\mathscr{L},\pm}) by:

  • UγN,±u(x)=γu(γ1x)U^{N,\pm}_{\gamma}u(x)=\gamma u(\gamma^{-1}x);

  • SγN,±u=eiϕγ|NuS_{\gamma}^{N,\pm}u=e^{i\phi_{\gamma}|_{N}}u;

  • TγN,±=UγNSγN,±T_{\gamma}^{N,\pm}=U_{\gamma}^{N}\circ S_{\gamma}^{N,\pm},

where uL2(E0,,±N)u\in L^{2}(E^{N}_{0,\mathscr{L},\pm}) and xNx\in N. Note that (2.1) and (2.2) continue to hold if we restrict ω\omega, η\eta, and ϕ\phi to NN and work with the de Rham differential on NN instead of MM. It follows that the operator DE0,,+ND^{E^{N}_{0,\mathscr{L},+}} is equivariant with respect the projective action TN,+T^{N,+}. Since Γ\Gamma acts on NN cocompactly, DE0,,+ND^{E^{N}_{0,\mathscr{L},+}} has a (Γ,σ)(\Gamma,\sigma)-invariant higher index

IndΓ,σ(DE0,,+N)K0(Cr(Γ,σ))\operatorname{Ind}_{\Gamma,\sigma}\big{(}D^{E^{N}_{0,\mathscr{L},+}}\big{)}\in K_{0}(C^{*}_{r}(\Gamma,\sigma))

by Definition 3.32 and (3.22). Equivalently, this index can be formulated as in (5.3), using a bounded transform on a Hilbert module. As mentioned previously, we will adopt the latter in order to follow more closely the exposition of [11].

With these preparations, we have the following:

Theorem 5.7 ((Γ,σ)(\Gamma,\sigma)-Callias-type index theorem).
IndΓ,σ(D+Φ)=IndΓ,σ(DE0,,+N)K0(Cr(Γ,σ)).\operatorname{Ind}_{\Gamma,\sigma}(D+\Phi)=\operatorname{Ind}_{\Gamma,\sigma}(D^{E^{N}_{0,\mathscr{L},+}})\in K_{0}(C^{*}_{r}(\Gamma,\sigma)). (5.10)

This is the projective analogue of the equivariant Callias-type index theorem [11, Theorem 3.4]. The proof is analogous to that in the untwisted setting, once we make the following modifications:

  1. (1)

    instead of the Hilbert C(G)C^{*}(G)-modules used in [11], we work with Hilbert Cr(Γ,σ)C^{*}_{r}(\Gamma,\sigma)-modules;

  2. (2)

    instead of the GG-equivariant bundle SS used in [11] (resp. bundles derived from SS), we work with the bundle EE_{\mathscr{L}} from subsection 5.2 (or bundles derived from it);

  3. (3)

    instead of GG-invariant differential operators, we work with (Γ,σ)(\Gamma,\sigma)-invariant operators, as constructed above;

  4. (4)

    instead of the index indexG\operatorname{index}_{G} from [11, section 3], we work with IndΓ,σ\operatorname{Ind}_{\Gamma,\sigma}.

Given these similarities, we will for the most part only sketch the proof of Theorem 5.7, and invite the reader who is interested in a more detailed discussion to [11, section 5]. Nevertheless, let us give a detailed example of how one of the key technical tools used in the proof of [11, Theorem 3.4] can be adapted to the projective setting, namely the relative index theorem for Callias-type operators [11, Theorem 4.13]. The twisted analogue of that theorem is as follows.

For j=1,2j=1,2, let MjM_{j}, EjE_{j}, DjD_{j} and Φj\Phi_{j} be as MM, EE, DD, and Φ\Phi were in Definition 5.2. Suppose there exist Γ\Gamma-invariant, cocompact hypersurfaces NjMjN_{j}\subseteq M_{j}, Γ\Gamma-invariant tubular neighbourhoods UjNjU_{j}\supseteq N_{j}, and a Γ\Gamma-equivariant isometry ψ:U1U2\psi\colon U_{1}\to U_{2} such that

  • ψ(N1)=N2\psi(N_{1})=N_{2};

  • ψ(E2|U2)E1|U1\psi^{*}(E_{2}|_{U_{2}})\cong E_{1}|_{U_{1}};

  • ψ(2|U2)=1|U1\psi^{*}(\nabla_{2}|_{U_{2}})=\nabla_{1}|_{U_{1}}, where j\nabla_{j} is the Clifford connection used to define DjD_{j};

  • Φ1|U1\Phi_{1}|_{U_{1}} corresponds to Φ2|U2\Phi_{2}|_{U_{2}} via ψ\psi.

Suppose that Mj=XjNjYjM_{j}=X_{j}\cup_{N_{j}}Y_{j} for closed, Γ\Gamma-invariant subsets Xj,YjMjX_{j},Y_{j}\subseteq M_{j}. We identify N1N_{1} and N2N_{2} via ψ\psi and simply write NN for this manifold. Construct

M3X1NY2;M4X2NY1.M_{3}\coloneqq X_{1}\cup_{N}Y_{2};\qquad M_{4}\coloneqq X_{2}\cup_{N}Y_{1}.

For j=3,4j=3,4, let EjE_{j}, DjD_{j} and Φj\Phi_{j} be obtained from the corresponding data on M1M_{1} and M2M_{2} by cutting and gluing along U1U2U_{1}\cong U_{2} via ψ\psi. For j=1,2,3,4j=1,2,3,4, form the Hilbert Cr(Γ,σ)C^{*}_{r}(\Gamma,\sigma)-modules jσ\mathcal{E}^{\sigma}_{j} as in Definition 5.4.

Proposition 5.8.

In the above situation,

IndΓ,σ(D1+Φ1)+IndΓ,σ(D2+Φ2)=IndΓ,σ(D3+Φ3)+IndΓ,σ(D4+Φ4)K0(Cr(Γ,σ)).\operatorname{Ind}_{\Gamma,\sigma}(D_{1}+\Phi_{1})+\operatorname{Ind}_{\Gamma,\sigma}(D_{2}+\Phi_{2})=\operatorname{Ind}_{\Gamma,\sigma}(D_{3}+\Phi_{3})+\operatorname{Ind}_{\Gamma,\sigma}(D_{4}+\Phi_{4})\in K_{0}(C^{*}_{r}(\Gamma,\sigma)).
Proof.

(Compare the proof of [11, Theorem 4.13].) Define

σ1σ2σ3σ,op4σ,op,\mathcal{E}^{\sigma}\coloneqq\mathcal{E}^{\sigma}_{1}\oplus\mathcal{E}^{\sigma}_{2}\oplus\mathcal{E}_{3}^{\sigma,\operatorname{op}}\oplus\mathcal{E}_{4}^{\sigma,\operatorname{op}},

where a superscript op\operatorname{op} indicates reversal of the 2\mathbb{Z}_{2}-grading on the given module. Similar to (5.3), define

Fj(Dj+Φj)((Dj+Φj)2+fj)1/2,F_{j}\coloneqq(D_{j}+\Phi_{j})\bigl{(}(D_{j}+\Phi_{j})^{2}+f_{j}\bigr{)}^{-1/2},

for j=1,2,3,4j=1,2,3,4, and

FF1F2F3F4.F\coloneqq F_{1}\oplus F_{2}\oplus F_{3}\oplus F_{4}.

For j=1,2j=1,2, let χXj,χYjC(Mj)\chi_{X_{j}},\chi_{Y_{j}}\in C^{\infty}(M_{j}) be real-valued functions such that:

  1. (i)

    supp(χXj)XjUj\operatorname{supp}(\chi_{X_{j}})\subseteq X_{j}\cup U_{j} and supp(χYj)YjUj\operatorname{supp}(\chi_{Y_{j}})\subseteq Y_{j}\cup U_{j};

  2. (ii)

    ψ(χX2|U2)=χX1|U1\psi^{*}(\chi_{X_{2}}|_{U_{2}})=\chi_{X_{1}}|_{U_{1}} and ψ(χY2|U2)=χY1|U1\psi^{*}(\chi_{Y_{2}}|_{U_{2}})=\chi_{Y_{1}}|_{U_{1}}; and

  3. (iii)

    χXj2+χYj2=1\chi_{X_{j}}^{2}+\chi_{Y_{j}}^{2}=1.

We view pointwise multiplication by these functions as operators

χX1:1σ3σ;χY1:1σ4σ;χY2:2σ3σ;χX2:2σ4σ.\begin{split}\chi_{X_{1}}\colon&\mathcal{E}_{1}^{\sigma}\to\mathcal{E}_{3}^{\sigma};\\ \chi_{Y_{1}}\colon&\mathcal{E}_{1}^{\sigma}\to\mathcal{E}_{4}^{\sigma};\end{split}\qquad\begin{split}\chi_{Y_{2}}\colon&\mathcal{E}_{2}^{\sigma}\to\mathcal{E}_{3}^{\sigma};\\ \chi_{X_{2}}\colon&\mathcal{E}_{2}^{\sigma}\to\mathcal{E}_{4}^{\sigma}.\end{split} (5.11)

Define the operator

Xγ(00χX1χY100χY2χX2χX1χY200χY1χX200)(σ),X\coloneqq\gamma\begin{pmatrix}0&0&-\chi_{X_{1}}^{*}&-\chi_{Y_{1}}^{*}\\ 0&0&-\chi_{Y_{2}}^{*}&\chi_{X_{2}}^{*}\\ \chi_{X_{1}}&\chi_{Y_{2}}&0&0\\ \chi_{Y_{1}}&-\chi_{X_{2}}&0&0\end{pmatrix}\in\mathcal{B}(\mathcal{E}^{\sigma}),

where γ\gamma is the grading operator on σ\mathcal{E}^{\sigma}. Then XX is an odd, self-adjoint operator on σ\mathcal{E}^{\sigma}. Further, using properties (ii) and (iii), one verifies directly that X2=1.X^{2}=1. Let l\mathbb{C}l denote the Clifford algebra generated by XX. It follows from a discussion analogous to [11, section 4] that

XF+FX𝒦(σ).XF+FX\in\mathcal{K}(\mathcal{E}^{\sigma}).

Since XX generates l\mathbb{C}l and anticommutes with FF modulo 𝒦(σ)\mathcal{K}(\mathcal{E}^{\sigma}), the pair (σ,F)(\mathcal{E}^{\sigma},F) is a Kasparov (l,Cr(Γ,σ))(\mathbb{C}l,C_{r}^{*}(\Gamma,\sigma))-cycle. Its class is mapped to [σ,F]K(Cr(Γ,σ))[\mathcal{E}^{\sigma},F]\in K_{*}(C^{*}_{r}(\Gamma,\sigma)) by the homomorphism induced by the pullback along the inclusion l\mathbb{C}\hookrightarrow\mathbb{C}l. By [4, Lemma 1.15], that homomorphism is zero. Hence

[σ,F]=0K0(Cr(Γ,σ)),[\mathcal{E}^{\sigma},F]=0\in K_{0}(C^{*}_{r}(\Gamma,\sigma)),

which is equivalent to the theorem. ∎

Proof of Theorem 5.7.

The first and most important step is to use Proposition 5.8 to reduce the computation of IndΓ,σ(D+Φ)\operatorname{Ind}_{\Gamma,\sigma}(D+\Phi) to the index of a (Γ,σ)(\Gamma,\sigma)-invariant Callias-type operator on the cylinder N×N\times\mathbb{R}. This is done by following the same geometric steps as in [11, subsection 5.3], only applied to the bundle EE instead of the bundle SS used there. Where [11, Theorem 4.13] was used, we now apply Proposition 5.8. In [11], a homotopy invariance property of equivariant Callias-type operators [11, Proposition 4.9] was proved, together with the fact that the index of a Callias-type operator does not change if one modifies the potential Φ\Phi on a cocompact subset [11, Corollary 4.10]. The proofs of both of these properties carry over to the projective setting after making the modifications (1) – (4) from above.

More precisely, the above discussion reduces the computation of IndΓ,σ(D+Φ)\operatorname{Ind}_{\Gamma,\sigma}(D+\Phi) to the (Γ,σ)(\Gamma,\sigma)-index of the operator (5.12) below. To define this operator, let E0,E_{0,\mathscr{L}} be as in (5.5), and denote by E0,NE^{N}_{0,\mathscr{L}} its restriction to NN. Let

E0,,±N×N×E^{N\times\mathbb{R}}_{0,\mathscr{L},\pm}\to N\times\mathbb{R}

be the pullbacks of E0,,±NNE^{N}_{0,\mathscr{L},\pm}\to N along the canonical projection N×NN\times\mathbb{R}\to N. Then E0,,±N×E^{N\times\mathbb{R}}_{0,\mathscr{L},\pm} are Clifford bundles over T(N×)T(N\times\mathbb{R}), with Clifford action

c^(v,t)=c(v+tn^),\widehat{c}(v,t)=c(v+t\widehat{n}),

where vTNv\in TN, tt\in\mathbb{R}, n^\widehat{n} is the normal vector field to NN pointing into M+M_{+}, and cc is Clifford multiplication on E0,E_{0,\mathscr{L}}. Let E0,,±N\nabla^{E^{N}_{0,\mathscr{L},\pm}} and DE0,,±ND^{E^{N}_{0,\mathscr{L},\pm}} be as in (5.8) and (5.9). By pulling back E0,,±N\nabla^{E^{N}_{0,\mathscr{L},\pm}} along N×N\times\mathbb{R}\to\mathbb{R} and composing with c^\widehat{c}, we obtain Dirac operators D0E0,,±N×D^{E^{N\times\mathbb{R}}_{0,\mathscr{L},\pm}}_{0} acting on sections of E0,,±N×E^{N\times\mathbb{R}}_{0,\mathscr{L},\pm}. In particular, operator D0E0,,+N×D^{E^{N\times\mathbb{R}}_{0,\mathscr{L},+}}_{0} is equivariant with respect to the pull-back of the projective action TN,+T^{N,+} to L2(E0,,+N×)L^{2}(E^{N\times\mathbb{R}}_{0,\mathscr{L},+}). Let χC()\chi\in C^{\infty}(\mathbb{R}) be an odd function such that χ(t)=t\chi(t)=t for all t2t\geq 2. Let χN×\chi_{N\times\mathbb{R}} be its pullback along the projection N×N\times\mathbb{R}\to\mathbb{R}. For our Callias-type operator, we will take two copies of the bundle E0,,+N×E^{N\times\mathbb{R}}_{0,\mathscr{L},+}. Define

DE0,,+N×=(0D0E0,,+N×D0E0,,+N×0),D^{E^{N\times\mathbb{R}}_{0,\mathscr{L},+}}=\begin{pmatrix}0&D^{E^{N\times\mathbb{R}}_{0,\mathscr{L},+}}_{0}\\ D^{E^{N\times\mathbb{R}}_{0,\mathscr{L},+}}_{0}&0\end{pmatrix},

acting on smooth sections of E0,,+N×E0,,+N×E^{N\times\mathbb{R}}_{0,\mathscr{L},+}\oplus E^{N\times\mathbb{R}}_{0,\mathscr{L},+}. Then the endomorphism

χN×=(0iχN×iχN×0)\chi^{N\times\mathbb{R}}=\begin{pmatrix}0&i\chi_{N\times\mathbb{R}}\\ -i\chi_{N\times\mathbb{R}}&0\end{pmatrix}

is admissible for DE0,,+N×D^{E^{N\times\mathbb{R}}_{0,\mathscr{L},+}} in the sense of Definition 5.2, and

DE0,,+N×+χN×D^{E^{N\times\mathbb{R}}_{0,\mathscr{L},+}}+\chi^{N\times\mathbb{R}} (5.12)

is a (Γ,σ)(\Gamma,\sigma)-invariant Callias-type operator on N×N\times\mathbb{R}. By the discussion in the first paragraph of this proof, we have

IndΓ,σ(D+Φ)=IndΓ,σ(DE0,,+N×+χN×).\operatorname{Ind}_{\Gamma,\sigma}(D+\Phi)=\operatorname{Ind}_{\Gamma,\sigma}(D^{E^{N\times\mathbb{R}}_{0,\mathscr{L},+}}+\chi^{N\times\mathbb{R}}). (5.13)

It then suffices to prove that

IndΓ,σ(DE0,,+N×+χN×)=IndΓ,σ(DE0,,+N),\operatorname{Ind}_{\Gamma,\sigma}(D^{E^{N\times\mathbb{R}}_{0,\mathscr{L},+}}+\chi^{N\times\mathbb{R}})=\operatorname{Ind}_{\Gamma,\sigma}(D^{E^{N}_{0,\mathscr{L},+}}), (5.14)

which is the projective analogue of [11, Proposition 5.7]. For this, note that the operator DE0,,+N×+χN×D^{E^{N\times\mathbb{R}}_{0,\mathscr{L},+}}+\chi^{N\times\mathbb{R}} can be written explicitly as

(0DE0,,+NDE0,,+N0)1C()+γE0,,+N(0iddtiddt0)+1C(E0,,+N)(0iχiχ0),\begin{pmatrix}0&D^{E^{N}_{0,\mathscr{L},+}}\\ D^{E^{N}_{0,\mathscr{L},+}}&0\end{pmatrix}\otimes 1_{C^{\infty}(\mathbb{R})}+\gamma_{E^{N}_{0,\mathscr{L},+}}\otimes\begin{pmatrix}0&i\frac{d}{dt}\\ i\frac{d}{dt}&0\end{pmatrix}+1_{C^{\infty}(E^{N}_{0,\mathscr{L},+})}\otimes\begin{pmatrix}0&-i\chi\\ i\chi&0\end{pmatrix},

where γE0,,+N\gamma_{E^{N}_{0,\mathscr{L},+}} is a grading on E0,,+NE^{N}_{0,\mathscr{L},+} defined as i-i times Clifford multiplication by the unit normal vector field on NN pointing into M+M_{+}. The equality (5.14) then follows from the fact that the kernel of iddt±iχi\frac{d}{dt}\pm i\chi in C()C^{\infty}(\mathbb{R}) is one-dimensional. Combining (5.13) and (5.14) concludes the proof. ∎

Remark 5.9.

Theorem 5.7 continues to hold if we replace σ\sigma by σs\sigma^{s} for any ss\in\mathbb{R}. In this case, the connection \nabla^{\mathscr{L}} from (5.4) would be replaced by ,s=d+isη\nabla^{\mathscr{L},s}=d+is\eta.

5.3. Proof of Theorem 1.7

Proof of Theorem 1.7.

This proof is similar to, but more subtle than, that of [11, Theorem 2.1], as it involves an additional scaling argument along with the use of an appropriate partition of unity. Hence we will give the full details.

First note that by a suspension argument, we only need to consider odd-dimensional MM. In this case, let 𝒮\mathcal{S} be the spinor bundle over MM, let ∂̸\not{\partial} be the spin-Dirac operator, and let 𝒮=𝒮\mathcal{S}_{\mathscr{L}}=\mathcal{S}\otimes\mathscr{L} for a Γ\Gamma-trivial line bundle \mathscr{L}. For ss\in\mathbb{R}, let ,s\nabla^{\mathscr{L},s} be the Hermitian connection on \mathscr{L} defined by the one-form isηis\eta. In the notation of subsection 5.2, take E0,=𝒮E_{0,\mathscr{L}}=\mathcal{S}_{\mathscr{L}} and D0sD_{0}^{s} be the Dirac operator associated to the connection 𝒮1+1,s\nabla^{\mathcal{S}}\otimes 1+1\otimes\nabla^{\mathscr{L},s}. Let

Ds=(0D0sD0s0)D^{s}=\begin{pmatrix}0&D_{0}^{s}\\ D_{0}^{s}&0\end{pmatrix}

act on sections of 𝒮𝒮\mathcal{S}_{\mathscr{L}}\oplus\mathcal{S}_{\mathscr{L}}. We now construct a potential Φ\Phi that is admissible for DsD^{s}, for all ss, in the sense of Definition 5.2.

Let HH be as in the statement of the theorem. Then MH=XYM\setminus H=X\cup Y for disjoint open subsets XX and YY. Pick a cocompact subset KK of MM such that HKX¯H\subseteq K\subseteq\overline{X} and κ>0\kappa>0 on KK, and the distance from XKX\setminus K to YY is positive. Pick a Γ\Gamma-invariant function χC(M)\chi\in C^{\infty}(M) such that χ\chi equals 11 on YY and 1-1 on XKX\setminus K. Let Φ0\Phi_{0} be the endomorphism of 𝒮\mathcal{S}_{\mathscr{L}} given by pointwise multiplication by χ\chi. Now define DsD^{s} and Φ\Phi according to (5.6) and (5.7) respectively. Define the endomorphism

Φ=(0iχiχ0).\Phi=\begin{pmatrix}0&i\chi\\ -i\chi&0\end{pmatrix}.

One finds that D0s=∂̸+isc(η)D_{0}^{s}=\not{\partial}+isc(\eta). Together with the fact that [isc(η),iχ]=0[isc(\eta),i\chi]=0, this implies that

{Ds,Φ}=i(sc(dχ)00sc(dχ)).\{D^{s},\Phi\}=-i\begin{pmatrix}sc(d\chi)&0\\ 0&-sc(d\chi)\end{pmatrix}.

By construction, the estimate Φ2{Ds,Φ}+1\Phi^{2}\geq\|\{D^{s},\Phi\}\|+1 holds pointwise on MKM\setminus K, hence Φ\Phi is admissible for DsD^{s}, for all ss\in\mathbb{R}, in the sense of Definition 5.2.

Next, in the notation of subsection 5.1, take M=KM_{-}=K. Then N=MN=\partial M_{-} is a disjoint union NHN_{-}\cup H for a cocompact subset NN_{-} such that f|N=1f|_{N_{-}}=-1. In this case,

E+N=𝒮|H.E^{N}_{+}=\mathcal{S}_{\mathscr{L}}|_{H}. (5.15)

By Theorem 5.7 and the proof of Theorem 1.5 (see also Remark 5.9), it suffices to show that IndΓ,σ(Ds+Φ)=0\operatorname{Ind}_{\Gamma,\sigma}(D^{s}+\Phi)=0 for all sufficiently small ss. For convenience, let us write Bλs=Ds+λΦB_{\lambda}^{s}=D^{s}+\lambda\Phi for λ>0\lambda>0. By a homotopy argument, IndΓ,σ(Bs)=IndΓ,σ(Bλs)\operatorname{Ind}_{\Gamma,\sigma}(B^{s})=\operatorname{Ind}_{\Gamma,\sigma}(B^{s}_{\lambda}) for any positive λ\lambda, so it suffices to show that

IndΓ,σ(Bλs)=0\operatorname{Ind}_{\Gamma,\sigma}(B^{s}_{\lambda})=0

for some λ>0\lambda>0 and all sufficiently small ss. Note that the endomorphism λΦ\lambda\Phi is still admissible for DsD^{s}, and (5.15) continues to hold. Let KK^{\prime} be an arbitrary cocompact neighbourhood of KK. By construction,

(Bλs)2λ2(B^{s}_{\lambda})^{2}\geq\lambda^{2} (5.16)

on MKM\setminus K. On the set KK^{\prime}, we can obtain an estimate as follows. Letting s\nabla^{s} be the connection used to define DsD^{s}, we have

(Bλs)2\displaystyle(B^{s}_{\lambda})^{2} =(Ds)2+{Ds,λΦ}+Φ2\displaystyle=(D^{s})^{2}+\{D^{s},\lambda\Phi\}+\Phi^{2}
=ss+κ4+isλc(ω)+{Ds,λΦ}+λ2Φ2\displaystyle={\nabla^{s}}^{*}\nabla^{s}+\frac{\kappa}{4}+is\lambda c(\omega)+\{D^{s},\lambda\Phi\}+\lambda^{2}\Phi^{2}
κ4+isλc(ω)+{Ds,λΦ}+λ2Φ2.\displaystyle\geq\frac{\kappa}{4}+is\lambda c(\omega)+\{D^{s},\lambda\Phi\}+\lambda^{2}\Phi^{2}. (5.17)

Let κ0=infxKκ(x)>0\kappa_{0}=\inf_{x\in K}\kappa(x)>0 by cocompactness of KK. Then

  • on KK: there exist s0,λ0>0s_{0},\lambda_{0}>0 such for all s<s0s<s_{0} and λ<λ0\lambda<\lambda_{0}, the endomorphism (5.3) is bounded below by κ08\frac{\kappa_{0}}{8};

  • on KKK^{\prime}\setminus K: since κ0\kappa\geq 0 and {Ds,λΦ}=0\{D^{s},\lambda\Phi\}=0, the endomorphism (5.3) is bounded below by isλc(ω)+λ2Φ2.is\lambda c(\omega)+\lambda^{2}\Phi^{2}. By cocompactness of KK^{\prime}, there exist s1,λ1s_{1},\lambda_{1} such that isλc(ω)+λ2Φ2κ08is\lambda c(\omega)+\lambda^{2}\Phi^{2}\geq\frac{\kappa_{0}}{8} for all s<s1s<s_{1} and λ<λ1\lambda<\lambda_{1}.

Combining this with (5.16), we see that the estimate

(Bλs)2κ08(B^{s}_{\lambda})^{2}\geq\frac{\kappa_{0}}{8}

holds on both KK^{\prime} and MKM\setminus K for all s<inf{s0,s1}s<\inf\{s_{0},s_{1}\} and λ<inf{λ0,λ1,κ08}\lambda<\inf\{\lambda_{0},\lambda_{1},\frac{\kappa_{0}}{8}\}. To combine these these estimates, let ϕ1,ϕ2\phi_{1},\phi_{2} smooth functions M[0,1]M\to[0,1] such that

  • {ϕ12,ϕ22}\{\phi_{1}^{2},\phi_{2}^{2}\} is a partition of unity on MM;

  • supp(ϕ1)K\operatorname{supp}(\phi_{1})\subseteq K^{\prime} and supp(ϕ2)MK\operatorname{supp}(\phi_{2})\subseteq M\setminus K.

For any ε>0\varepsilon>0, we may take KK^{\prime} to be sufficiently large so that dϕ1,dϕ2<ε\|d\phi_{1}\|_{\infty},\|d\phi_{2}\|_{\infty}<\varepsilon. Since BsB^{s} is a perturbation of ∂̸\not{\partial} by an endomorphism,

[ϕi,Bs]=[ϕi,∂̸]=c(dϕi),[\phi_{i},B^{s}]=[\phi_{i},\not{\partial}]=c(d\phi_{i}), (5.18)

for i=1,2i=1,2, hence ϕiBλs=Bλsϕi+[ϕi,∂̸]\phi_{i}B_{\lambda}^{s}=B_{\lambda}^{s}\phi_{i}+[\phi_{i},\not{\partial}]. For any uL2(𝒮)u\in L^{2}(\mathcal{S}_{\mathscr{L}}), one computes that

ϕiBλsu,ϕiBλsu\displaystyle\langle\phi_{i}B_{\lambda}^{s}u,\phi_{i}B_{\lambda}^{s}u\rangle =[ϕi,∂̸]u,ϕiBλsu+Bλsϕiu,ϕiBλsu\displaystyle=\langle[\phi_{i},\not{\partial}]u,\phi_{i}B_{\lambda}^{s}u\rangle+\langle B_{\lambda}^{s}\phi_{i}u,\phi_{i}B_{\lambda}^{s}u\rangle
=[ϕi,∂̸]u,ϕiBλsu+ϕiBλsu,[ϕi,∂̸]u\displaystyle=\langle[\phi_{i},\not{\partial}]u,\phi_{i}B_{\lambda}^{s}u\rangle+\langle\phi_{i}B_{\lambda}^{s}u,[\phi_{i},\not{\partial}]u\rangle
+[∂̸,ϕi]u,[ϕi,∂̸]u+Bλsϕiu,Bλsϕiu\displaystyle\qquad\qquad+\langle[\not{\partial},\phi_{i}]u,[\phi_{i},\not{\partial}]u\rangle+\langle B_{\lambda}^{s}\phi_{i}u,B_{\lambda}^{s}\phi_{i}u\rangle
Bλsϕiu22εBλsuuε2u2,\displaystyle\geq\|B_{\lambda}^{s}\phi_{i}u\|^{2}-2\varepsilon\|B_{\lambda}^{s}u\|\|u\|-\varepsilon^{2}\|u\|^{2},

where the norms and inner products are taken in L2(𝒮)L^{2}(\mathcal{S}_{\mathscr{L}}) and we have used that ϕi1\|\phi_{i}\|_{\infty}\leq 1. It follows that

(Bλs)2u,u\displaystyle\langle(B_{\lambda}^{s})^{2}u,u\rangle =ϕ12Bλsu,Bλsu+ϕ22Bλsu,Bλsu\displaystyle=\langle\phi_{1}^{2}B_{\lambda}^{s}u,B_{\lambda}^{s}u\rangle+\langle\phi_{2}^{2}B_{\lambda}^{s}u,B_{\lambda}^{s}u\rangle
2(Bλsϕiu22εBλsuuε2u2)\displaystyle\geq 2(\|B_{\lambda}^{s}\phi_{i}u\|^{2}-2\varepsilon\|B_{\lambda}^{s}u\|\|u\|-\varepsilon^{2}\|u\|^{2})
(κ04ε2)u24εBλsuu.\displaystyle\geq\big{(}\frac{\kappa_{0}}{4}-\varepsilon^{2}\big{)}\|u\|^{2}-4\varepsilon\|B_{\lambda}^{s}u\|\|u\|.

Hence (Bλsu+2εu)2(κ04+3ε2)u2,(\|B_{\lambda}^{s}u\|+2\varepsilon\|u\|)^{2}\geq\big{(}\frac{\kappa_{0}}{4}+3\varepsilon^{2}\big{)}\|u\|^{2}, so that

Bλsu(κ04+3ε22ε)u.\|B_{\lambda}^{s}u\|\geq\Big{(}\sqrt{\frac{\kappa_{0}}{4}+3\varepsilon^{2}}-2\varepsilon\Big{)}\|u\|.

Taking ε\varepsilon small enough, and hence KK^{\prime} large enough, we see that (Bλs)2(B_{\lambda}^{s})^{2} is strictly positive. Thus BλsB^{s}_{\lambda} is invertible for λ<inf{λ0,λ1,κ08}\lambda<\inf\{\lambda_{0},\lambda_{1},\frac{\kappa_{0}}{8}\} and s<inf{s0,s1}s<\inf\{s_{0},s_{1}\}, whence IndΓ,σ(Bλs)=0\operatorname{Ind}_{\Gamma,\sigma}(B^{s}_{\lambda})=0. ∎


6. A quantitative obstruction in the non-cocompact setting

When M/ΓM/\Gamma is non-compact, we can use quantitative KK-theory to give obstructions to the existence of Γ\Gamma-invariant metrics of positive scalar curvature on MM. This uses the fact that the twisted Roe algebra is naturally filtered by propagation, making it an example of a geometric CC^{*}-algebra. We now review these concepts.

6.1. Geometric CC^{*}-algebras and quantitative K-theory

Definition 6.1.

A unital CC^{*}-algebra AA is geometric if it admits a filtration {Ar}r>0\{A_{r}\}_{r>0} satisfying the following properties:

  1. (i)

    ArArA_{r}\subseteq A_{r^{\prime}} if rrr\leq r^{\prime};

  2. (ii)

    ArArAr+rA_{r}A_{r^{\prime}}\subseteq A_{r+r^{\prime}};

  3. (iii)

    r=0Ar\bigcup_{r=0}^{\infty}A_{r} is dense in AA.

If AA is non-unital, then its unitization A+A^{+}, viewed as AA\oplus\mathbb{C} as as a vector space, is a geometric CC^{*}-algebra with filtration {Ar}r>0.\{A_{r}\oplus\mathbb{C}\}_{r>0}. In addition, for each nn, the matrix algebra Mn(A)M_{n}(A) is a geometric CC^{*}-algebra with filtration {Mn(Ar)}r>0.\{M_{n}(A_{r})\}_{r>0}.

Definition 6.2 ([5, Definition 2.15]).

Let AA be a geometric CC^{*}-algebra. For 0<ε<1200<\varepsilon<\frac{1}{20}, r>0r>0, and N1N\geq 1,

  • an element eAe\in A is called an (ε,r,N)(\varepsilon,r,N)-quasiidempotent if

    e2e<ε,eAr,max(e,1A+e)N;\|e^{2}-e\|<\varepsilon,\qquad e\in A_{r},\qquad\max(\|e\|,\|1_{A^{+}}-e\|)\leq N;
  • if AA is unital, an element uAu\in A is called an (ε,r,N)(\varepsilon,r,N)-quasiinvertible if uAru\in A_{r}, uN\|u\|\leq N, and there exists vArv\in A_{r} with

    vN,max(uv1,vu1)<ε.\|v\|\leq N,\qquad\max(\|uv-1\|,\|vu-1\|)<\varepsilon.

    The pair (u,v)(u,v) is called an (ε,r,N)(\varepsilon,r,N)-quasiinverse pair.

The quantitative KK-groups K0ε,r,N(A)K_{0}^{\varepsilon,r,N}(A) and K1ε,r,N(A)K_{1}^{\varepsilon,r,N}(A) are defined by collecting together all quasiidempotents and quasiinvertibles over all matrix algebras, quotienting by an equivalence relation, and taking the Gröthendieck completion.

Definition 6.3 ([5, subsection 3.1]).

Let AA be a unital geometric CC^{*}-algebra. Let r>0r>0, 0<ε<1200<\varepsilon<\frac{1}{20}, and N>0N>0.

  1. (i)

    Denote by Idemε,r,N(A)\textnormal{Idem}^{\varepsilon,r,N}(A) the set of (ε,r,N)(\varepsilon,r,N)-quasiidempotents in AA. For each positive integer nn, let

    Idemnε,r,N(A)=Idemε,r,N(Mn(A)).\textnormal{Idem}_{n}^{\varepsilon,r,N}(A)=\textnormal{Idem}^{\varepsilon,r,N}(M_{n}(A)).

    We have inclusions Idemnε,r,N(A)Idemn+1ε,r,N(A)\textnormal{Idem}_{n}^{\varepsilon,r,N}(A)\hookrightarrow\textnormal{Idem}_{n+1}^{\varepsilon,r,N}(A) given by e(e000).e\mapsto\begin{pmatrix}e&0\\ 0&0\end{pmatrix}. Set

    Idemε,r,N(A)=n=1Idemnε,r,N(A).\textnormal{Idem}_{\infty}^{\varepsilon,r,N}(A)=\bigcup_{n=1}^{\infty}\textnormal{Idem}_{n}^{\varepsilon,r,N}(A).

    Define an equivalence relation \sim on Idemε,r,N(A)\textnormal{Idem}_{\infty}^{\varepsilon,r,N}(A) by efe\sim f if ee and ff are (4ε,r,4N)(4\varepsilon,r,4N)-homotopic in M(A)M_{\infty}(A). Denote the equivalence class of an element eIdemε,r,N(A)e\in\textnormal{Idem}_{\infty}^{\varepsilon,r,N}(A) by [e][e]. Define addition on Idemε,r,N(A)/\textnormal{Idem}_{\infty}^{\varepsilon,r,N}(A)/\sim by

    [e]+[f]=[e00f].[e]+[f]=\begin{bmatrix}e&0\\ 0&f\end{bmatrix}.

    With this operation, Idemε,r,N(A)/\textnormal{Idem}_{\infty}^{\varepsilon,r,N}(A)/\sim is an abelian monoid with identity [0][0]. Let K0ε,r,N(A)K_{0}^{\varepsilon,r,N}(A) denote its Grothendieck completion.

  2. (ii)

    Denote by GLε,r,N(A)GL^{\varepsilon,r,N}(A) the set of (ε,r,N)(\varepsilon,r,N)-quasiinvertibles in AA. For each positive integer nn, let

    GLnε,r,N(A)=GLnε,r,N(Mn(A)).GL_{n}^{\varepsilon,r,N}(A)=GL_{n}^{\varepsilon,r,N}(M_{n}(A)).

    We have inclusions GLnε,r,N(A)GLn+1ε,r,N(A)GL_{n}^{\varepsilon,r,N}(A)\hookrightarrow GL_{n+1}^{\varepsilon,r,N}(A) given by u(u001).u\mapsto\begin{pmatrix}u&0\\ 0&1\end{pmatrix}. Set

    GLε,r,N(A)=n=1GLnε,r,N(A).GL_{\infty}^{\varepsilon,r,N}(A)=\bigcup_{n=1}^{\infty}GL_{n}^{\varepsilon,r,N}(A).

    Define an equivalence relation \sim on GLε,r,N(A)GL_{\infty}^{\varepsilon,r,N}(A) by efe\sim f if uu and vv are (4ε,2r,4N)(4\varepsilon,2r,4N)-homotopic in M(A)M_{\infty}(A). Denote the equivalence class of an element uGLε,r,N(A)u\in GL_{\infty}^{\varepsilon,r,N}(A) by [u][u]. Define addition on GLε,r,N(A)/GL_{\infty}^{\varepsilon,r,N}(A)/\sim by

    [u]+[v]=[u00v].[u]+[v]=\begin{bmatrix}u&0\\ 0&v\end{bmatrix}.

    With this operation, GLε,r,N(A)/GL_{\infty}^{\varepsilon,r,N}(A)/\sim is an abelian group [1][1].

Remark 6.4.

If AA is a non-unital geometric CC^{*}-algebra, then we have a canonical *-homomorphism π:A+\pi\colon A^{+}\to\mathbb{C}. Using contractivity of π\pi, we have homomorphisms

π:Kiε,r,N(A+)Kiε,r,N(),\pi_{*}\colon K_{i}^{\varepsilon,r,N}(A^{+})\to K_{i}^{\varepsilon,r,N}(\mathbb{C}),

where i=0i=0 or 11. Define Kiε,r,N(A)=ker(π)K_{i}^{\varepsilon,r,N}(A)=\ker(\pi_{*}).

The following result on quasiidempotents and quasiinvertibles is useful.

Lemma 6.5 ([12, Lemma 3.4]).

Let AA be a geometric CC^{*}-algebra. If ee is an (ε,r,N)(\varepsilon,r,N)-idempotent in AA, and fArf\in A_{r} satisfies

fN,ef<εe2e2N+1,\|f\|\leq N,\qquad\|e-f\|<\frac{\varepsilon-\|e^{2}-e\|}{2N+1},

then ff is a quasiidempotent that is (ε,r,N)(\varepsilon,r,N)-homotopic to ee. In particular, if

f<ε2N+1,\|f\|<\frac{\varepsilon}{2N+1},

then the class of ff is zero in K0ε,r,N(A)K_{0}^{\varepsilon,r,N}(A).

Suppose that AA is unital and (u,v)(u,v) is an (ε,r,N)(\varepsilon,r,N)-quasiinverse pair in AA. If aAra\in A_{r} satisfies

aN,ua<εmax(uv1,vu1)N,\|a\|\leq N,\qquad\|u-a\|<\frac{\varepsilon-\max(\|uv-1\|,\|vu-1\|)}{N},

then aa is a quasiinvertible that is (ε,r,N)(\varepsilon,r,N)-homotopic to uu. In particular, if

1a<εN,\|1-a\|<\frac{\varepsilon}{N},

then the class of aa is zero in K1ε,r,N(A)K_{1}^{\varepsilon,r,N}(A).

There is a homomorphism of abelian groups

Ψ:Kε,r,N(A)K(A),\Psi\colon K_{*}^{\varepsilon,r,N}(A)\to K_{*}(A), (6.1)

preserving the 2\mathbb{Z}_{2}-grading, where Kε,r,N(A)K_{*}^{\varepsilon,r,N}(A) (resp. K(A)K_{*}(A)) denotes the direct sum of the quantitative (resp. operator) K0K_{0} and K1K_{1}-groups.

6.2. The quantitative higher index

Fix 0<ε<1200<\varepsilon<\tfrac{1}{20} and N7N\geq 7. For each ss\in\mathbb{R}, let the algebras [M;L2(𝒮)]Γ,σs\mathbb{C}[M;L^{2}(\mathcal{S}_{\mathscr{L}})]^{\Gamma,\sigma^{s}} and C(M;L2(𝒮))Γ,σsC^{*}(M;L^{2}(\mathcal{S}_{\mathscr{L}}))^{\Gamma,\sigma^{s}} be as in Definition 3.27, with σ\sigma and the projective representation TT replaced by σs\sigma^{s} and TsT^{s}, as in Definition 2.3.

For each r>0r>0, define the subspace

[M;L2(𝒮)]rΓ,σs{T[M;L2(𝒮)]Γ,σs:prop(T)r}\mathbb{C}[M;L^{2}(\mathcal{S}_{\mathscr{L}})]^{\Gamma,\sigma^{s}}_{r}\coloneqq\{T\in\mathbb{C}[M;L^{2}(\mathcal{S}_{\mathscr{L}})]^{\Gamma,\sigma^{s}}\colon\textnormal{prop}(T)\leq r\} (6.2)

of C(M;L2(𝒮))Γ,σsC^{*}(M;L^{2}(\mathcal{S}_{\mathscr{L}}))^{\Gamma,\sigma^{s}}. Then with respect to the filtration

{[M;L2(𝒮)]Γ,σs}r>0,\{\mathbb{C}[M;L^{2}(\mathcal{S}_{\mathscr{L}})]^{\Gamma,\sigma^{s}}\}_{r>0},

C(M;L2(𝒮))Γ,σsC^{*}(M;L^{2}(\mathcal{S}_{\mathscr{L}}))^{\Gamma,\sigma^{s}} is a geometric CC^{*}-algebra in the sense of Definition 6.1. As in [12], this structure allows us define a refinement of the higher index that takes values in the quantitative KK-groups of C(M;L2(𝒮))Γ,σsC^{*}(M;L^{2}(\mathcal{S}_{\mathscr{L}}))^{\Gamma,\sigma^{s}}. The construction is similar to that in [12, subsection 3.2.2], so we will be brief.

The (Γ,σs)(\Gamma,\sigma^{s})-invariant higher index IndΓ,σs(D)\operatorname{Ind}_{\Gamma,\sigma^{s}}(D) given in Definition 3.32 can be represented explicitly as follows. Let χ\chi be a normalising function. If dimM\dim M is even, define the idempotent

pχ(D)=([(1χ(D)2)2]1,1[χ(D)(1χ(D)2)]1,2[χ(D)(2χ(D)2)(1χ(D)2)]2,1[χ(D)2(2χ(D)2)]2,2),p_{\chi}(D)=\begin{pmatrix}\left[(1-\chi(D)^{2})^{2}\right]_{1,1}&\,\,\left[\chi(D)(1-\chi(D)^{2})\right]_{1,2}\\[4.30554pt] \left[\chi(D)(2-\chi(D)^{2})(1-\chi(D)^{2})\right]_{2,1}&\,\,\left[\chi(D)^{2}(2-\chi(D)^{2})\right]_{2,2}\end{pmatrix}, (6.3)

where the notation [X]i,j[X]_{i,j} means the (i,j)(i,j)-th entry of the matrix XX. Then IndΓ,σ(D)\operatorname{Ind}_{\Gamma,\sigma}(D) is represented by the difference of idempotents

Aχ(D)=pχ(D)(0001).A_{\chi}(D)=p_{\chi}(D)-\begin{pmatrix}0&0\\ 0&1\end{pmatrix}. (6.4)

For dimM\dim M odd, IndΓ,σ(D)\operatorname{Ind}_{\Gamma,\sigma}(D) can be represented by the unitary

Aχ(D)=eπi(χ+1)(D).A_{\chi}(D)=e^{\pi i(\chi+1)}(D). (6.5)

Even-dimensional MM

Choose a normalizing function χ\chi such that

suppχ^[r5,r5].\operatorname{supp}\widehat{\chi}\subseteq\left[-\frac{r}{5},\frac{r}{5}\right]. (6.6)

Let Aχ(Ds)=pχ(Ds)(0001)A_{\chi}(D^{s})=p_{\chi}(D^{s})-\begin{pmatrix}0&0\\ 0&1\end{pmatrix} as in (6.4). Then the (ε,r,N)(\varepsilon,r,N)-quantitative maximal higher index of DD is the class

IndΓ,σs,L2ε,r,N(Ds)=[pχ(Ds)][0001]K0ε,r,N(C(M;L2(𝒮))Γ,σs).\operatorname{Ind}_{\Gamma,\sigma^{s},L^{2}}^{\varepsilon,r,N}(D^{s})=\left[p_{\chi}(D^{s})\right]-\begin{bmatrix}0&0\\ 0&1\end{bmatrix}\in K_{0}^{\varepsilon,r,N}(C^{*}(M;L^{2}(\mathcal{S}_{\mathscr{L}}))^{\Gamma,\sigma^{s}}).

Odd-dimensional MM

For each integer n0n\geq 0 define polynomials

fn(x)\displaystyle f_{n}(x) =k=0n(2πix)kk!,gn(x)=fn(x)(k=1n(2πi)kk!)x2.\displaystyle=\sum_{k=0}^{n}\frac{(2\pi ix)^{k}}{k!},\qquad g_{n}(x)=f_{n}(x)-\left(\sum_{k=1}^{n}\frac{(2\pi i)^{k}}{k!}\right)x^{2}. (6.7)

One finds that as nn\to\infty, the difference e2πixgn(x)e^{2\pi ix}-g_{n}(x) converges uniformly to 0 for xx in the interval [2,2][-2,2]. Let m=m(ε,N)m=m(\varepsilon,N) be the smallest number such that

|gm(x)gm(x)1|<ε,\displaystyle|g_{m}(x)g_{m}(-x)-1|<\varepsilon,
|e2πixgm(x)|<1,\displaystyle|e^{2\pi ix}-g_{m}(x)|<1, (6.8)

for all x[2,2]x\in[-2,2]. Pick a normalizing function χ\chi satisfying

suppχ^[rdeggm,rdeggm].\operatorname{supp}\widehat{\chi}\subseteq\left[-\frac{r}{\deg g_{m}},\frac{r}{\deg g_{m}}\right]. (6.9)

Then the operator

Sχ=χ(Ds)+12S_{\chi}=\frac{\chi(D^{s})+1}{2}

has propagation at most rdeggm\frac{r}{\deg g_{m}} and spectrum contained in [12,32][-\frac{1}{2},\frac{3}{2}]. Then gm(Sχ(Ds))g_{m}(S_{\chi}(D^{s})) is an (ε,r,N)(\varepsilon,r,N)-quasiinvertible in the unitisation of C(M;L2(𝒮))Γ,σsC^{*}(M;L^{2}(\mathcal{S}_{\mathscr{L}}))^{\Gamma,\sigma^{s}}. It was shown in [12, subsection 3.2.2] that

IndΓ,σs(Ds)=[Aχ(Ds)]=[gm(Sχ)]K1(C(M;L2(𝒮))Γ,σs).\operatorname{Ind}_{{\Gamma,\sigma^{s}}}(D^{s})=[A_{\chi}(D^{s})]=[g_{m}(S_{\chi})]\in K_{1}(C^{*}(M;L^{2}(\mathcal{S}_{\mathscr{L}}))^{\Gamma,\sigma^{s}}).

The (ε,r,N)(\varepsilon,r,N)-quantitative higher index of DsD^{s} is the class

IndΓ,σsε,r,N(Ds)=[gm(Sχ)]K1ε,r,N(C(M;L2(𝒮))Γ,σs).\operatorname{Ind}_{{\Gamma,\sigma^{s}}}^{\varepsilon,r,N}(D^{s})=[g_{m}(S_{\chi})]\in K_{1}^{\varepsilon,r,N}(C^{*}(M;L^{2}(\mathcal{S}_{\mathscr{L}}))^{\Gamma,\sigma^{s}}).
Remark 6.6.

First, although in the above constructions we needed to make a choice of χ\chi, the quantitative higher index obtained is independent of this choice.

Second, the (Γ,σ)(\Gamma,\sigma)-higher index of DsD^{s} relates to its quantitative refinement by IndΓ(Ds)=Ψ(IndΓ(Ds))\operatorname{Ind}_{\Gamma}(D^{s})=\Psi(\operatorname{Ind}_{\Gamma}(D^{s})), where Ψ\Psi is the homomorphism from (6.1).

6.3. A quantitative obstruction

We now prove Theorem 1.9. This uses the construction of the twisted higher index from subsection 2.3 in terms of twisted Roe algebras, which are geometric CC^{*}-algebras in the sense of [21]. The result we obtain generalises [12, Theorem 1.1].

Proof.

The technique of the proof is as in [12, section 4]. The differences are that we now work with the reduced rather than the maximal version of the twisted Roe algebra, and that bounds on κ\kappa used in that paper are now replaced by bounds on the endomorphism κ+4isc(ω)\kappa+4isc(\omega). By Lemma 4.7, we have

(Ds)2=ss+κ4+isc(ω).(D^{s})^{2}=\nabla^{s*}\nabla^{s}+\frac{\kappa}{4}+isc(\omega).

Suppose that κ+4isc(ω)Cs\kappa+4isc(\omega)\geq C_{s} holds as an estimate on operators on L2(𝒮)L^{2}(\mathcal{S}_{\mathscr{L}}). Let χ\chi be a normalizing function whose distributional Fourier transform χ^\widehat{\chi} is supported on some finite interval [s,s][-s,s] for s>0s>0. For each t>0t>0, let χt\chi_{t} be the normalizing function defined by

χt(u)=χ(tu),\chi_{t}(u)=\chi(tu), (6.10)

uu\in\mathbb{R}. Let Aχ(Ds)A_{\chi}(D^{s}) be the index representative defined using χ\chi.

If MM is even-dimensional, let

Aχ(u)((1χ(u)2)2χ(t)(1χ(u)2)χ(u)(2χ(u)2)(1χ(u)2)χ(u)2(2χ(u)2)1),u.A_{\chi}(u)\coloneqq\begin{pmatrix}(1-\chi(u)^{2})^{2}&\,\,\chi(t)(1-\chi(u)^{2})\\[4.30554pt] \chi(u)(2-\chi(u)^{2})(1-\chi(u)^{2})&\,\,\chi(u)^{2}(2-\chi(u)^{2})-1\end{pmatrix},\quad u\in\mathbb{R}.

Let u0>0u_{0}>0 and a function α\alpha be such that

Aχ(u)<ε2N+1\|A_{\chi}(u)\|<\frac{\varepsilon}{2N+1} (6.11)

whenever |1χ(u)2|<α(ε)|1-\chi(u)^{2}|<\alpha(\varepsilon) for all uu such that |u|>u0|u|>u_{0}, where the norm of Aχ(u)A_{\chi}(u) is taken in M2()M_{2}(\mathbb{C}). Note that for N7N\geq 7, (6.11) also implies that Aχ2(u)Aχ(u)<ε\|A_{\chi}^{2}(u)-A_{\chi}(u)\|<\varepsilon if |u|>u0|u|>u_{0}. By (6.10), we have

|1χ2u0Cs(u)2|=|1χ(2u0uCs)2|<α(ε)\Big{|}1-\chi_{\frac{2u_{0}}{\sqrt{C_{s}}}}(u)^{2}\Big{|}=\Big{|}1-\chi\left(\tfrac{2u_{0}u}{\sqrt{C_{s}}}\right)^{2}\Big{|}<\alpha(\varepsilon) (6.12)

whenever u(Cs2,Cs2)u\in\mathbb{R}\setminus(-\frac{\sqrt{C_{s}}}{2},\frac{\sqrt{C_{s}}}{2}), while

supp(χ^2u0Cs(Ds))[2u0Css,2u0Css].\operatorname{supp}\Big{(}\widehat{\chi}_{\tfrac{2u_{0}}{\sqrt{C_{s}}}}(D^{s})\Big{)}\subseteq\left[-\tfrac{2u_{0}}{\sqrt{C_{s}}}s,\tfrac{2u_{0}}{\sqrt{C_{s}}}s\right].

It follows that Aχ2u0Cs(Ds)A_{\chi_{\frac{2u_{0}}{\sqrt{C_{s}}}}}(D^{s}) is an (ε,10u0Css,N)(\varepsilon,\frac{10u_{0}}{\sqrt{C_{s}}}s,N)-quasiidempotent in 2×22\times 2-matrices over the unitisation of C(M;L2(𝒮))Γ,σsC^{*}(M;L^{2}(\mathcal{S}_{\mathscr{L}}))^{\Gamma,\sigma^{s}} with norm strictly less than ε2N+1\frac{\varepsilon}{2N+1}. By Lemma 6.5,

IndΓ,σs,L2ε,10u0Css,N(Ds)=0K0ε,10u0Css,N(C(M;L2(𝒮))Γ,σs).\displaystyle\operatorname{Ind}_{{\Gamma,\sigma^{s},L^{2}}}^{\varepsilon,\frac{10u_{0}}{\sqrt{C_{s}}}s,N}(D^{s})=0\in K_{0}^{\varepsilon,\frac{10u_{0}}{\sqrt{C_{s}}}s,N}(C^{*}(M;L^{2}(\mathcal{S}_{\mathscr{L}}))^{\Gamma,\sigma^{s}}).

Letting λ0=10u0s\lambda_{0}=10u_{0}s, we obtain IndΓ,σs,L2ε,λ0Cs,N(Ds)=0\operatorname{Ind}_{\Gamma,\sigma^{s},L^{2}}^{\varepsilon,\frac{\lambda_{0}}{\sqrt{C_{s}}},N}(D^{s})=0. Note that for any rλ0Csr\geq\frac{\lambda_{0}}{\sqrt{C_{s}}}, IndΓ,σs,L2ε,r,N(Ds)\operatorname{Ind}_{\Gamma,\sigma^{s},L^{2}}^{\varepsilon,r,N}(D^{s}) can also be represented by Aχ2u0Cs(Ds)A_{\chi_{\frac{2u_{0}}{\sqrt{C_{s}}}}}(D^{s}). The homomorphism

K0ε,λ0Cs,N(C(M;L2(𝒮))Γ,σs)K0ε,r,N(C(M;L2(𝒮))Γ,σs)K_{0}^{\varepsilon,\frac{\lambda_{0}}{\sqrt{C_{s}}},N}(C^{*}(M;L^{2}(\mathcal{S}_{\mathscr{L}}))^{\Gamma,\sigma^{s}})\to K_{0}^{\varepsilon,r,N}(C^{*}(M;L^{2}(\mathcal{S}_{\mathscr{L}}))^{\Gamma,\sigma^{s}})

induced by the inclusion

Idemε,λ0Cs,N(C(M;L2(𝒮))Γ,σs)Idemε,r,N(C(M;L2(𝒮))Γ,σs)\textnormal{Idem}_{\infty}^{\varepsilon,\frac{\lambda_{0}}{\sqrt{C_{s}}},N}(C^{*}(M;L^{2}(\mathcal{S}_{\mathscr{L}}))^{\Gamma,\sigma^{s}})\hookrightarrow\textnormal{Idem}_{\infty}^{\varepsilon,r,N}(C^{*}(M;L^{2}(\mathcal{S}_{\mathscr{L}}))^{\Gamma,\sigma^{s}})

takes IndΓ,σs,L2ε,λ0Cs,N(Ds)\operatorname{Ind}_{{\Gamma,\sigma^{s},L^{2}}}^{\varepsilon,\frac{\lambda_{0}}{\sqrt{C_{s}}},N}(D^{s}) to IndΓ,σs,L2ε,r,N(Ds)\operatorname{Ind}_{{\Gamma,\sigma^{s},L^{2}}}^{\varepsilon,r,N}(D^{s}), which therefore vanishes.

If MM is odd-dimensional, let m=m(ε,N)m=m(\varepsilon,N) and the polynomial gmg_{m} be as (6.2). Let χ\chi be a normalizing function satisfying (6.9), and let s=rdeggms=\frac{r}{\deg g_{m}}. Let u0>0u_{0}>0 be such that

1gm(Pχ(u))<εN\|1-g_{m}(P_{\chi}(u))\|<\frac{\varepsilon}{N}

whenever |1χ(u)2|<α(ε)|1-\chi(u)^{2}|<\alpha(\varepsilon) holds for all uu such that |u|>u0|u|>u_{0} or, equivalently, whenever

|1χ2u0Cs(u)2|=|1χ(2u0uCs)2|<α(ε)\Big{|}1-\chi_{\frac{2u_{0}}{\sqrt{C_{s}}}}(u)^{2}\Big{|}=\Big{|}1-\chi\left(\tfrac{2u_{0}u}{\sqrt{C_{s}}}\right)^{2}\Big{|}<\alpha(\varepsilon) (6.13)

for all u(Cs2,Cs2)u\in\mathbb{R}\setminus(-\frac{\sqrt{C_{s}}}{2},\frac{\sqrt{C_{s}}}{2}). Meanwhile,

supp(χ^2u0Cs(Ds))[2u0Css,2u0Css].\operatorname{supp}\Big{(}\widehat{\chi}_{\tfrac{2u_{0}}{\sqrt{C_{s}}}}(D^{s})\Big{)}\subseteq\left[-\tfrac{2u_{0}}{\sqrt{C_{s}}}s,\tfrac{2u_{0}}{\sqrt{C_{s}}}s\right].

Thus gm(Pχ2u0Cs(Ds))g_{m}(P_{\chi_{\frac{2u_{0}}{\sqrt{C_{s}}}}}(D^{s})) is an (ε,2mu0Css,N)(\varepsilon,\frac{2mu_{0}}{\sqrt{C_{s}}}s,N)-quasiinvertible in 2×22\times 2-matrices over the unitisation of C(M;L2(𝒮))Γ,σsC^{*}(M;L^{2}(\mathcal{S}_{\mathscr{L}}))^{\Gamma,\sigma^{s}} satisfying

1gm(Pχ2u0Cs(Ds))<εN.\Big{\|}1-g_{m}(P_{\chi_{\frac{2u_{0}}{\sqrt{C_{s}}}}}(D^{s}))\Big{\|}<\frac{\varepsilon}{N}. (6.14)

By Lemma 6.5,

IndΓ,σs,L2ε,2mu0Css,N(Ds)=0K1ε,2mu0Css,N(C(M;L2(𝒮))Γ,σs).\displaystyle\operatorname{Ind}_{{\Gamma,\sigma^{s},L^{2}}}^{\varepsilon,\frac{2mu_{0}}{\sqrt{C_{s}}}s,N}(D^{s})=0\in K_{1}^{\varepsilon,\frac{2mu_{0}}{\sqrt{C_{s}}}s,N}(C^{*}(M;L^{2}(\mathcal{S}_{\mathscr{L}}))^{\Gamma,\sigma^{s}}).

Letting λ0=2mu0s\lambda_{0}=2mu_{0}s, we obtain IndΓ,σs,L2ε,λ0Cs,N(Ds)=0\operatorname{Ind}_{{\Gamma,\sigma^{s}},L^{2}}^{\varepsilon,\frac{\lambda_{0}}{\sqrt{C_{s}}},N}(D^{s})=0. For any rλ0Csr\geq\frac{\lambda_{0}}{\sqrt{C_{s}}}, the element IndΓ,σs,L2ε,r,N(Ds)\operatorname{Ind}_{{\Gamma,\sigma^{s},L^{2}}}^{\varepsilon,r,N}(D^{s}) can also be represented by gm(Pχ2u0Cs(Ds))g_{m}(P_{\chi_{\frac{2u_{0}}{\sqrt{C_{s}}}}}(D^{s})). The homomorphism

K1ε,λ0c,N(C(M;L2(𝒮))Γ,σs)K1ε,r,N(C(M;L2(𝒮))Γ,σs)K_{1}^{\varepsilon,\frac{\lambda_{0}}{\sqrt{c}},N}(C^{*}(M;L^{2}(\mathcal{S}_{\mathscr{L}}))^{\Gamma,\sigma^{s}})\to K_{1}^{\varepsilon,r,N}(C^{*}(M;L^{2}(\mathcal{S}_{\mathscr{L}}))^{\Gamma,\sigma^{s}})

induced by the inclusion

GLε,λ0Cs,N((C(M;L2(𝒮))Γ,σs)+)GLε,r,N((C(M;L2(𝒮))Γ,σs)+)GL_{\infty}^{\varepsilon,\frac{\lambda_{0}}{\sqrt{C_{s}}},N}((C^{*}(M;L^{2}(\mathcal{S}_{\mathscr{L}}))^{\Gamma,\sigma^{s}})^{+})\hookrightarrow GL_{\infty}^{\varepsilon,r,N}((C^{*}(M;L^{2}(\mathcal{S}_{\mathscr{L}}))^{\Gamma,\sigma^{s}})^{+})

takes IndΓ,σs,L2ε,λ0Cs,N(Ds)\operatorname{Ind}_{{\Gamma,\sigma^{s},L^{2}}}^{\varepsilon,\frac{\lambda_{0}}{\sqrt{C_{s}}},N}(D^{s}) to IndΓ,σs,L2ε,r,N(Ds)\operatorname{Ind}_{{\Gamma,\sigma^{s},L^{2}}}^{\varepsilon,r,N}(D^{s}), which therefore vanishes. ∎


References

  • [1] Alejandro Adem, Johann Leida, and Yongbin Ruan. Orbifolds and stringy topology, volume 171 of Cambridge Tracts in Mathematics. Cambridge University Press, Cambridge, 2007.
  • [2] Paul Baum, Alain Connes, and Nigel Higson. Classifying space for proper actions and KK-theory of group CC^{\ast}-algebras. In CC^{\ast}-algebras: 1943–1993 (San Antonio, TX, 1993), volume 167 of Contemp. Math., pages 240–291. Amer. Math. Soc., Providence, RI, 1994.
  • [3] Nicole Berline, Ezra Getzler, and Michèle Vergne. Heat kernels and Dirac operators. Grundlehren Text Editions. Springer-Verlag, Berlin, 2004. Corrected reprint of the 1992 original.
  • [4] Ulrich Bunke. A KK-theoretic relative index theorem and Callias-type Dirac operators. Math. Ann., 303(2):241–279, 1995.
  • [5] Yeong Chyuan Chung. Quantitative KK-theory for Banach algebras. J. Funct. Anal., 274(1):278–340, 2018.
  • [6] A. Connes, M. Gromov, and H. Moscovici. Group cohomology with Lipschitz control and higher signatures. Geom. Funct. Anal., 3(1):1–78, 1993.
  • [7] Alain Connes. Noncommutative geometry. Academic Press, Inc., San Diego, CA, 1994.
  • [8] Alain Connes and Henri Moscovici. Cyclic cohomology, the Novikov conjecture and hyperbolic groups. Topology, 29(3):345–388, 1990.
  • [9] Mikhael Gromov and H. Blaine Lawson, Jr. Positive scalar curvature and the Dirac operator on complete Riemannian manifolds. Inst. Hautes Études Sci. Publ. Math., (58):83–196 (1984), 1983.
  • [10] Hao Guo. Index of equivariant Callias-type operators and invariant metrics of positive scalar curvature. J. Geom. Anal., 31(1):1–34, 2021.
  • [11] Hao Guo, Peter Hochs, and Varghese Mathai. Positive scalar curvature and an equivariant Callias-type index theorem for proper actions. Ann. K-theory, 6(2):319–356, 2021.
  • [12] Hao Guo, Zhizhang Xie, and Guoliang Yu. Quantitative KK-theory, positive scalar curvature, and band width. In Perspectives in Scalar Curvature (Eds. M. Gromov and H. B. Lawson). World Sci., to appear.
  • [13] Nigel Higson and John Roe. Analytic KK-homology. Oxford Mathematical Monographs. Oxford University Press, Oxford, 2000. Oxford Science Publications.
  • [14] Peter Hochs and Hang Wang. A fixed point formula and Harish-Chandra’s character formula. Proc. Lond. Math. Soc. (3), 116(1):1–32, 2018.
  • [15] Ronghui Ji. Smooth dense subalgebras of reduced group CC^{*}-algebras, Schwartz cohomology of groups, and cyclic cohomology. J. Funct. Anal., 107(1):1–33, 1992.
  • [16] Y. Kordyukov, V. Mathai, and M. Shubin. Equivalence of spectral projections in semiclassical limit and a vanishing theorem for higher traces in KK-theory. J. Reine Angew. Math., 581:193–236, 2005.
  • [17] Varghese Mathai. KK-theory of twisted group CC^{*}-algebras and positive scalar curvature. In Tel Aviv Topology Conference: Rothenberg Festschrift (1998), volume 231 of Contemp. Math., pages 203–225. Amer. Math. Soc., Providence, RI, 1999.
  • [18] Varghese Mathai. Heat kernels and the range of the trace on completions of twisted group algebras. In The ubiquitous heat kernel, volume 398 of Contemp. Math., pages 321–345. Amer. Math. Soc., Providence, RI, 2006. With an appendix by Indira Chatterji.
  • [19] H. Moscovici and F.-B. Wu. Localization of topological Pontryagin classes via finite propagation speed. Geom. Funct. Anal., 4(1):52–92, 1994.
  • [20] J. M. Osterburg and D. S. Passman. Trace methods in twisted group algebras. II. Proc. Amer. Math. Soc., 130(12):3495–3506, 2002.
  • [21] Hervé Oyono-Oyono and Guoliang Yu. On quantitative operator KK-theory. Ann. Inst. Fourier (Grenoble), 65(2):605–674, 2015.
  • [22] Jonathan Rosenberg. CC^{\ast}-algebras, positive scalar curvature, and the Novikov conjecture. III. Topology, 25(3):319–336, 1986.
  • [23] Larry B. Schweitzer. A short proof that Mn(A)M_{n}(A) is local if AA is local and Fréchet. Internat. J. Math., 3(4):581–589, 1992.
  • [24] Bai-Ling Wang and Hang Wang. Localized index and L2L^{2}-Lefschetz fixed-point formula for orbifolds. J. Differential Geom., 102(2):285–349, 2016.
  • [25] Hang Wang. L2L^{2}-index formula for proper cocompact group actions. J. Noncommut. Geom., 8(2):393–432, 2014.
  • [26] Rufus Willett and Guoliang Yu. Higher Index Theory. Cambridge Studies in Advanced Mathematics. Cambridge University Press, 2020.
  • [27] Zhizhang Xie and Guoliang Yu. Delocalized Eta Invariants, Algebraicity, and K-Theory of Group C*-Algebras. Int. Math. Res. Not. IMRN, 08 2019. rnz170.