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

Classical harmonic analysis viewed through the prism of noncommutative geometry

Cédric Arhancet
Abstract

The aim of this paper is to bridge noncommutative geometry with classical harmonic analysis on Banach spaces, focusing primarily on both classical and noncommutative Lp\mathrm{L}^{p} spaces. Introducing a notion of Banach Fredholm module, we define new abelian groups, K0(𝒜,)\mathrm{K}^{0}(\mathcal{A},\mathscr{B}) and K1(𝒜,)\mathrm{K}^{1}(\mathcal{A},\mathscr{B}), of K\mathrm{K}-homology associated with an algebra 𝒜\mathcal{A} and a suitable class \mathscr{B} of Banach spaces, such as the class of Lp\mathrm{L}^{p}-spaces. We establish index pairings of these groups with the K\mathrm{K}-theory groups of the algebra 𝒜\mathcal{A}. Subsequently, by considering (noncommutative) Hardy spaces, we uncover the natural emergence of Hilbert transforms, leading to Banach Fredholm modules and culminating in index theorems. Moreover, by associating each reasonable sub-Markovian semigroup of operators with a ¡¡Banach noncommutative manifold¿¿, we explain how this leads to (possibly kernel-degenerate) Banach Fredholm modules, thereby revealing the role of vectorial Riesz transforms in this context. Overall, our approach significantly integrates the analysis of operators on Lp\mathrm{L}^{p}-spaces into the expansive framework of noncommutative geometry, offering new perspectives.

00footnotetext: 2020 Mathematics subject classification: 58B34, 47D03, 46L80, 47B90
Key words: K\mathrm{K}-homology, K\mathrm{K}-theory, spectral triples, Riesz transforms, Hilbert transforms, Lp\mathrm{L}^{p}-spaces, Fredholm modules, noncommutative geometry.

1 Introduction

The aim of this paper is to provide a coherent framework that encompasses various aspects of harmonic analysis on Lp\mathrm{L}^{p}-spaces within the context of noncommutative geometry. We seek to elucidate connections that appear sporadically throughout the literature, clarifying these overlaps and coincidences. It is crucial to emphasize that noncommutative geometry is not merely a ¡¡generalization¿¿ of classical geometry. While it subsumes known spaces as particular cases, it offers a radically different approach to some classical spaces, such as fractals or leaf spaces of foliations, by introducing powerful analytical points of view and tools. This work can be seen as the next step in our ongoing research program, which began in our previous studies [ArK22], [Arh24a], [Arh24b], and [Arh24c].

In noncommutative geometry, the starting point is an algebra 𝒜\mathcal{A} (which may be commutative or not), representing the space, with its elements acting as bounded operators on a complex Hilbert space HH via a representation π:𝒜B(H)\pi\colon\mathcal{A}\to\mathrm{B}(H). This algebra 𝒜\mathcal{A} replaces the algebra C(K)\mathrm{C}(K) of continuous functions on a classical compact space KK endowed with a finite Borel measure, which acts on the complex Hilbert space L2(K)\mathrm{L}^{2}(K) by multiplication operators.

A fundamental concept in this framework is that of a Fredholm module. If 𝒜\mathcal{A} is unital, a Fredholm module over 𝒜\mathcal{A} is a bounded operator F:HHF\colon H\to H satisfying the three relations F2=IdHF^{2}=\mathrm{Id}_{H}, F=FF=F^{*} and [F,π(a)]=0[F,\pi(a)]=0 for any a𝒜a\in\mathcal{A} modulo a compact operator. If DD is the Dirac operator on a compact Riemannian spin manifold MM, the Fredholm module signD\operatorname{\mathrm{sign}}D canonically associated to MM encodes the conformal structure of the manifold, as discussed in [Bar07, Theorem 3.1 p. 388]. These operators enable the definition of K\mathrm{K}-homology groups K0(𝒜)\mathrm{K}^{0}(\mathcal{A}) and K1(𝒜)\mathrm{K}^{1}(\mathcal{A}), which are linked to the K\mathrm{K}-theory groups K0(𝒜)\mathrm{K}_{0}(\mathcal{A}) and K1(𝒜)\mathrm{K}_{1}(\mathcal{A}) of the algebra 𝒜\mathcal{A} through a pairing that leads to index theorems. The culmination of this theory is perhaps the index theorem of Connes-Moscovici itself [CoM95], [Hig02], and its generalization to the locally compact case [CGRS14], drawing inspiration from the Atiyah-Singer index theorem, whose applications are thoroughly covered in [BlB13]. For more information on noncommutative geometry, we refer to the books [GVF01], [EcI18] and the survey article [CPR11], as well as specific applications to solid state physics in [PSB16] and related works on K\mathrm{K}-theory and K\mathrm{K}-homology [Pus11], [Had03], [Had04], [NeT11], [EmN18], [FGMR19], [Ger22] all of which are grounded in the classical text [HiR00].

In this paper, we define a notion of Banach Fredholm module F:XXF\colon X\to X on an arbitrary Banach space XX over an algebra 𝒜\mathcal{A}, where elements act on the space XX via a representation π:𝒜B(X)\pi\colon\mathcal{A}\to\mathrm{B}(X). In Connes’ foundational work [Con94], an illustrative example of a Fredholm module is given by F=defiF\overset{\mathrm{def}}{=}\mathrm{i}\mathcal{H} where :L2()L2()\mathcal{H}\colon\mathrm{L}^{2}(\mathbb{R})\to\mathrm{L}^{2}(\mathbb{R}) is the Hilbert transform. This example is sometimes seen as an ¡¡exotic¿¿ Fredholm module. Here, we show that various Hilbert transforms on Lp\mathrm{L}^{p}-spaces satisfying F2=IdXF^{2}=\mathrm{Id}_{X} modulo a compact operator naturally arise as examples. Actually, we introduce huge classes of Banach Fredholm modules over group C\mathrm{C}^{*}-algebras, which have not been previously considered even in the Hilbertian case. If Mf:Lp(n)Lp(n)M_{f}\colon\mathrm{L}^{p}(\mathbb{R}^{n})\to\mathrm{L}^{p}(\mathbb{R}^{n}), gfgg\mapsto fg is the multiplication operator by a function ff, the commutators [,Mf][\mathcal{H},M_{f}] of the Hilbert transform \mathcal{H} (with n=1n=1) and more generally commutators [j,Mf][\mathcal{R}_{j},M_{f}] of Riesz transforms j:Lp(n)Lp(n)\mathcal{R}_{j}\colon\mathrm{L}^{p}(\mathbb{R}^{n})\to\mathrm{L}^{p}(\mathbb{R}^{n}) where 1jn1\leqslant j\leqslant n, on Lp\mathrm{L}^{p}-spaces or even similar operators in other contexts is a classical topic in analysis, see the survey [Wic20] and references therein, connected to Hardy spaces, spaces of functions of bounded/vanishing mean oscillation and factorization of functions, and we demonstrate that they also emerge naturally in the commutators of Banach Fredholm modules. Such a commutator is given by [F,π(a)][F,\pi(a)] for some a𝒜a\in\mathcal{A}. In this framework, we can introduce the Connes quantized differential

(1.1) ¯da=defi[F,π(a)],a𝒜\,{\raisebox{-0.56905pt}{$\mathchar 22\relax$}\mkern-12.0mu\mathrm{d}}a\overset{\mathrm{def}}{=}\mathrm{i}[F,\pi(a)],\quad a\in\mathcal{A}

of aa. In the case F=iF=\mathrm{i}\mathcal{H}, we have

(1.2) ([F,Ma]f)(x)=p.v.a(x)a(y)xyf(y)dy.\big{(}[F,M_{a}]f\big{)}(x)=\operatorname{p.v.}\int_{-\infty}^{\infty}\frac{a(x)-a(y)}{x-y}f(y)\mathop{}\mathopen{}\mathrm{d}y.

A well-known phenomenon in the Hilbertian context is the interplay between the differentiability of aa and the ¡¡degree of compactness¿¿ of the quantized differential ¯da\,{\raisebox{-0.56905pt}{$\mathchar 22\relax$}\mkern-12.0mu\mathrm{d}}a. In the Banach space context, we use some suitable quasi-Banach ideal Sappq(X)S^{q}_{\mathrm{app}}(X), where 0<q<0<q<\infty, relying on the concept of ss-numbers introduced by Pietsch, as a substitute of the Schatten space Sq(H)=def{T:HH:Tr|T|q<}S^{q}(H)\overset{\mathrm{def}}{=}\{T\colon H\to H:\operatorname{Tr}|T|^{q}<\infty\}, allowing us to define a notion of finitely summable Banach Fredholm module. The notion of ss-numbers is a generalization of the notion of singular value of operators acting on Hilbert spaces.

We equally introduce K\mathrm{K}-homology groups K0(𝒜,)\mathrm{K}^{0}(\mathcal{A},\mathscr{B}) and K1(𝒜,)\mathrm{K}^{1}(\mathcal{A},\mathscr{B}) associated to the algebra 𝒜\mathcal{A} and with a class \mathscr{B} of Banach spaces, as the class of (noncommutative) Lp\mathrm{L}^{p}-spaces. We explore their pairings with the K\mathrm{K}-theory groups K0(𝒜)\mathrm{K}_{0}(\mathcal{A}) and K1(𝒜)\mathrm{K}_{1}(\mathcal{A}) of the algebra 𝒜\mathcal{A}. In the simplest case, the pairing gives the index theorem of Gohberg-Krein for a Toeplitz operator Ta:Hp(𝕋)Hp(𝕋)T_{a}\colon\mathrm{H}^{p}(\mathbb{T})\to\mathrm{H}^{p}(\mathbb{T}) acting on the classical Hardy space Hp(𝕋)\mathrm{H}^{p}(\mathbb{T}) on the circle for any 1<p<1<p<\infty, generalizing the well-known case p=2p=2.

Connes demonstrated how to associate a canonical cyclic cocycle, known as the Chern character, with a finitely summable Fredholm module, and how this character can be used to compute the index pairing between the K\mathrm{K}-theory of 𝒜\mathcal{A} and the K\mathrm{K}-homology class of the Fredholm module. We show that our framework admits a similar Chern character. The (odd) Chern character in K-homology is defined from the quantized calculus (1.1) associated to a Banach Fredholm module by

ChnF(a0,a1,,an)=defcnTr(F¯da0¯da1¯dan),a0,a1,,an𝒜,\mathrm{Ch}_{n}^{F}(a_{0},a_{1},\dots,a_{n})\overset{\mathrm{def}}{=}c_{n}\operatorname{Tr}(F\,{\raisebox{-0.56905pt}{$\mathchar 22\relax$}\mkern-12.0mu\mathrm{d}}a_{0}\,{\raisebox{-0.56905pt}{$\mathchar 22\relax$}\mkern-12.0mu\mathrm{d}}a_{1}\cdots\,{\raisebox{-0.56905pt}{$\mathchar 22\relax$}\mkern-12.0mu\mathrm{d}}a_{n}),\qquad a_{0},a_{1},\dots,a_{n}\in\mathcal{A},

for some sufficiently large odd integer nn and where cnc_{n} is some suitable constant. Here, we use the trace Tr\operatorname{Tr} on the space Sapp1(X)S^{1}_{\mathrm{app}}(X) which is the unique continuous extension of the trace defined on the space of finite-rank operators acting on the Banach space XX.

Within the Hilbertian setting of Connes, it is well-known that a Fredholm module can be constructed from a compact spectral triple (i.e. a noncommutative compact Riemannian spin manifold). Such a compact spectral triple (𝒜,H,D)(\mathcal{A},H,D) consists of a selfadjoint operator DD, defined on a dense subspace of the Hilbert space HH, and satisfying certain axioms. Specifically, D1D^{-1} must be compact on KerD\operatorname{Ker}D^{\perp}, often referred to as the ¡¡unit length¿¿ or ¡¡line element¿¿, denoted by ds\mathop{}\mathopen{}\mathrm{d}s since we can introduce the noncommutative integral \int with a Dixmier trace.

The Dirac operator DD on a spin Riemannian compact manifold MM can be used for constructing a classical example of spectral triple. Several examples in different contexts are discussed in the survey [CoM08]. The Fredholm module obtained from a spectral triple (𝒜,H,D)(\mathcal{A},H,D) is built using the bounded operator

(1.3) signD:HH,\operatorname{\mathrm{sign}}D\colon H\to H,

constructed with the spectral theorem.

In this paper, we replace Connes’ spectral triples by the Banach compact spectral triples of [ArK22]. Such a triple (𝒜,X,D)(\mathcal{A},X,D) is composed of a Banach space XX, a representation π:𝒜B(X)\pi\colon\mathcal{A}\to\mathrm{B}(X) and a bisectorial operator DD on a Banach space XX that admits a bounded H(Σθbi)\mathrm{H}^{\infty}(\Sigma_{\theta}^{\mathrm{bi}}) functional calculus on the open bisector Σθbi=defΣθ(Σθ)\Sigma_{\theta}^{\mathrm{bi}}\overset{\mathrm{def}}{=}\Sigma_{\theta}\cup(-\Sigma_{\theta}) where Σθ=def{z\{0}:|argz|<θ}\Sigma_{\theta}\overset{\mathrm{def}}{=}\big{\{}z\in\mathbb{C}\backslash\{0\}:\>|\arg z|<\theta\big{\}}, with 0<θ<π20<\theta<\frac{\pi}{2}. We refer to Definition 7.4 for a precise definition with an assumption on some commutators. Roughly speaking, this means that the spectrum σ(D)\sigma(D) of DD is a subset of the closed bisector Σωbi¯\overline{\Sigma^{\mathrm{bi}}_{\omega}} for some ω(0,θ)\omega\in(0,\theta) as in Figure 2, that we have an appropriate ¡¡resolvent estimate¿¿ and that

(1.4) f(D)XXfH(Σθbi)\left\|f(D)\right\|_{X\to X}\lesssim\left\|f\right\|_{\mathrm{H}^{\infty}(\Sigma^{\mathrm{bi}}_{\theta})}

for any suitable function ff of the algebra H(Σθbi)\mathrm{H}^{\infty}(\Sigma^{\mathrm{bi}}_{\theta}) of all bounded holomorphic functions defined on the bisector Σσbi\Sigma^{\mathrm{bi}}_{\sigma}. Here ¡¡suitable¿¿ means regularly decaying at 0 and at \infty. In broad terms, the operator f(D)f(D) is defined by a ¡¡Cauchy integral¿¿

(1.5) f(D)=12πiΣνbif(z)R(z,D)dzf(D)=\frac{1}{2\pi\mathrm{i}}\int_{\partial\Sigma^{\mathrm{bi}}_{\nu}}f(z)R(z,D)\mathop{}\mathopen{}\mathrm{d}z

by integrating over the boundary of a larger bisector Σνbi\Sigma^{\mathrm{bi}}_{\nu} using the resolvent operator R(z,D)=def(zD)1R(z,D)\overset{\mathrm{def}}{=}(z-D)^{-1}, where ω<ν<θ\omega<\nu<\theta. Note that the boundedness of such a functional calculus is not free, contrary to the case of the functional calculus of a selfadjoint operator. We refer to our paper [Arh24c] for concrete examples of operator with such a bounded functional calculus, where we use a notion of curvature for obtaining it. Using the function sign\operatorname{\mathrm{sign}} defined by sign(z)=def1Σθ(z)1Σθ(z)\operatorname{\mathrm{sign}}(z)\overset{\mathrm{def}}{=}1_{\Sigma_{\theta}}(z)-1_{-\Sigma_{\theta}}(z), we show as the hilbertian case that signD:XX\operatorname{\mathrm{sign}}D\colon X\to X is a Fredholm module. This notion of functional calculus was popularized in the paper [AKM06], which contains a (second) solution to famous Kato’s square root problem solved in [AHLMT02] and in [AKM06] (see also [HLM02] and [Tch01]).

Let us explain how this approach integrates with ours, starting with the simplest case, the one-dimensional scenario. Consider a function aL()a\in\mathrm{L}^{\infty}(\mathbb{R}) such that Rea(x)κ>0\operatorname{Re}a(x)\geqslant\kappa>0 for almost all xx\in\mathbb{R} and the multiplication operator Ma:L2()L2()M_{a}\colon\mathrm{L}^{2}(\mathbb{R})\to\mathrm{L}^{2}(\mathbb{R}), faff\mapsto af. Following the approach of [AKM06], we can consider the unbounded operator

(1.6) D=def[0ddxMaddx0],D\overset{\mathrm{def}}{=}\begin{bmatrix}0&-\frac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}x}M_{a}\\ \frac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}x}&0\end{bmatrix},

acting on the complex Hilbert space L2()L2()\mathrm{L}^{2}(\mathbb{R})\oplus\mathrm{L}^{2}(\mathbb{R}), where ddxMa\frac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}x}M_{a} denotes the composition ddxMa\frac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}x}\circ M_{a}. Note that the operator DD is not selfadjoint in general. However, by [AKM06, Theorem 3.1 (i) p. 465], DD admits a bounded H(Σθbi)\mathrm{H}^{\infty}(\Sigma_{\theta}^{\mathrm{bi}}) functional calculus for some angle 0<θ<π20<\theta<\frac{\pi}{2}. Using D2=[ddxMaddx00d2dx2Ma]D^{2}=\begin{bmatrix}-\frac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}x}M_{a}\frac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}x}&0\\ 0&-\frac{\mathop{}\mathopen{}\mathrm{d}^{2}}{\mathop{}\mathopen{}\mathrm{d}x^{2}}M_{a}\end{bmatrix}, we see that formally we have a bounded operator

(1.7) signD=D(D2)12=[0ddx(ddxMaddx)120].\operatorname{\mathrm{sign}}D=D(D^{2})^{-\frac{1}{2}}=\begin{bmatrix}0&*\\ \frac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}x}(-\frac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}x}M_{a}\frac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}x})^{-\frac{1}{2}}&0\end{bmatrix}.

It is immediate to obtain the obtain the estimate dfdxL2()(ddxMaddx)12fL2()\left\|\frac{\mathop{}\mathopen{}\mathrm{d}f}{\mathop{}\mathopen{}\mathrm{d}x}\right\|_{\mathrm{L}^{2}(\mathbb{R})}\lesssim\left\|\big{(}-\frac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}x}M_{a}\frac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}x}\big{)}^{\frac{1}{2}}f\right\|_{\mathrm{L}^{2}(\mathbb{R})}. Actually, a slightly more elaborate argument gives the Kato square root estimate in one dimension

(1.8) (ddxMaddx)12fL2()dfdxL2(),fW1,2().\left\|\bigg{(}-\frac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}x}M_{a}\frac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}x}\bigg{)}^{\frac{1}{2}}f\right\|_{\mathrm{L}^{2}(\mathbb{R})}\approx\left\|\frac{\mathop{}\mathopen{}\mathrm{d}f}{\mathop{}\mathopen{}\mathrm{d}x}\right\|_{\mathrm{L}^{2}(\mathbb{R})},\quad f\in\mathrm{W}^{1,2}(\mathbb{R}).

Using the homomorphism π:Cc()B(L2()L2())\pi\colon\mathrm{C}_{c}^{\infty}(\mathbb{R})\mapsto\mathrm{B}(\mathrm{L}^{2}(\mathbb{R})\oplus\mathrm{L}^{2}(\mathbb{R})), aMaMaa\mapsto M_{a}\oplus M_{a}. We will easily show that (Cc(),L2()L2(),D)(\mathrm{C}_{c}^{\infty}(\mathbb{R}),\mathrm{L}^{2}(\mathbb{R})\oplus\mathrm{L}^{2}(\mathbb{R}),D) is a Banach locally compact spectral triple, which is not a locally compact spectral triple in the classical hilbertian sense.

It is important to realize that all the theory of sub-Markovian semigroups acting on Lp\mathrm{L}^{p}-spaces can be integrated into the notion of Banach spectral triples. Indeed the L2\mathrm{L}^{2}-generator A2-A_{2} of such semigroup (Tt)t0(T_{t})_{t\geqslant 0} acting on the Hilbert space L2(Ω)\mathrm{L}^{2}(\Omega) can be written A2=A_{2}=\partial^{*}\partial where \partial is a (unbounded) closed derivation defined on a dense subspace of L2(Ω)\mathrm{L}^{2}(\Omega) with values in a Hilbert L(Ω)\mathrm{L}^{\infty}(\Omega)-bimodule \mathcal{H}. Here Tt=etA2T_{t}=\mathrm{e}^{-tA_{2}} for any t0t\geqslant 0. The map \partial can be seen as an ¡¡abstract¿¿ analogue of the gradient operator \nabla of a smooth Riemannian manifold MM, which is a closed operator defined on a subspace of L2(M)\mathrm{L}^{2}(M) into the space L2(M,TM)\mathrm{L}^{2}(M,\mathrm{T}M) satisfying the relation Δ=-\Delta=\nabla^{*}\nabla where Δ\Delta is the Laplace-Beltrami operator and where =div\nabla^{*}=-\mathrm{div}.

This fundamental result allows anyone to introduce a triple (L(Ω),L2(Ω)2,D)(\mathrm{L}^{\infty}(\Omega),\mathrm{L}^{2}(\Omega)\oplus_{2}\mathcal{H},D) associated to the semigroup in the spirit of the previous Banach spectral triples. Here DD is the unbounded selfadjoint operator acting on a dense subspace of the Hilbert space L2(Ω)2\mathrm{L}^{2}(\Omega)\oplus_{2}\mathcal{H} defined by

(1.9) D=def[00].D\overset{\mathrm{def}}{=}\begin{bmatrix}0&\partial^{*}\\ \partial&0\end{bmatrix}.

It is possible in this context to introduce a homomorphism π:L(Ω)B(L2(Ω)2)\pi\colon\mathrm{L}^{\infty}(\Omega)\to\mathrm{B}(\mathrm{L}^{2}(\Omega)\oplus_{2}\mathcal{H}), see (11.8). Now, suppose that 1<p<1<p<\infty. Sometimes, the map \partial induces a closable unbounded operator :domLp()𝒳p\partial\colon\operatorname{dom}\partial\subset\mathrm{L}^{p}(\mathcal{M})\to\mathcal{X}_{p} for some Banach space 𝒳p\mathcal{X}_{p}. So we can consider the Lp\mathrm{L}^{p}-realization of the previous operator DD as acting on a dense subspace of the Banach space Lp(Ω)p𝒳p\mathrm{L}^{p}(\Omega)\oplus_{p}\mathcal{X}_{p}. If DD is bisectorial and admits a bounded H(Σθbi)\mathrm{H}^{\infty}(\Sigma_{\theta}^{\mathrm{bi}}) functional calculus then using the equalities D2=[00]D^{2}=\begin{bmatrix}\partial^{*}\partial&0\\ 0&\partial\partial^{*}\end{bmatrix} and A2=A_{2}=\partial^{*}\partial, we see that formally we have a bounded operator

(1.10) signD=D(D2)12=[0A120].\operatorname{\mathrm{sign}}D=D(D^{2})^{-\frac{1}{2}}=\begin{bmatrix}0&*\\ \partial A^{-\frac{1}{2}}&0\end{bmatrix}.

From this, it is apparent that the vectorial Riesz transform A12:Lp(Ω)𝒳p\partial A^{-\frac{1}{2}}\colon\mathrm{L}^{p}(\Omega)\to\mathcal{X}_{p} is bounded and appears in the operator F=defsignDF\overset{\mathrm{def}}{=}\operatorname{\mathrm{sign}}D and particularly with the possible pairing with the group K0(𝒜)\mathrm{K}_{0}(\mathcal{A}) of K\mathrm{K}-theory for a suitable subalgebra 𝒜\mathcal{A} of the algebra L(Ω)\mathrm{L}^{\infty}(\Omega), as we will see. Of course, we can consider more generally sub-Markovian semigroups acting on the noncommutative Lp\mathrm{L}^{p}-space Lp()\mathrm{L}^{p}(\mathcal{M}) of a von Neumann algebra \mathcal{M} in the previous discussion.

Structure of the Paper

This paper is structured as follows. Section 2 provides the necessary background and revisits key notations, as well as essential results required for our work. In Section 3, we introduce the concept of (odd or even) Banach Fredholm modules and define new groups, K0(𝒜,)\mathrm{K}^{0}(\mathcal{A},\mathscr{B}) and K1(𝒜,)\mathrm{K}^{1}(\mathcal{A},\mathscr{B}), in K\mathrm{K}-homology, associated with an algebra 𝒜\mathcal{A} and a suitable class \mathscr{B} of Banach spaces. Section 4 focuses on establishing pairings with the K\mathrm{K}-theory groups of the algebra 𝒜\mathcal{A}, which is central to our approach. In Section 5, we introduce a suitable notion of summability for Banach Fredholm modules, revisiting key concepts related to approximation numbers of operators acting on Banach spaces. Section 6 is dedicated to the Chern character associated with finitely summable Banach Fredholm modules. Following the classical Hilbert space approach, we demonstrate how the pairing between K\mathrm{K}-homology and K\mathrm{K}-theory groups can be described using the Chern character. In Section 7, we revisit the notion of Banach spectral triples and explain how Banach Fredholm modules can be constructed from these triples. Section 8 explores concrete examples of triples constructed using Dirac operators, leading to Banach Fredholm modules, where we compute the pairings with K\mathrm{K}-theory in specific cases. We also reveal the connection between the summability of Banach spectral triples and Sobolev embedding theorems. In Section 9, we present examples of Banach Fredholm modules in the context of noncommutative Hardy spaces, leading to new Banach Fredholm modules on group C\mathrm{C}^{*}-algebras, even within the Hilbert space framework. Section 10 present a Banach Fredholm module on the reduced C\mathrm{C}^{*}-algebra of the free group 𝔽\mathbb{F}_{\infty}. Finally, in Section 11 discusses how to associate a Banach spectral triple with each reasonable sub-Markovian semigroup. We also observe that the corresponding Banach Fredholm module, which may be kernel-degenerate, is closely linked to the vectorial Riesz transform associated to the semigroup.

2 Preliminaries

K\mathrm{K}-theory

We refer to the books [Bl98], [GVF01], [RLL00] and [WeO93] for more information. Two idempotents e,fe,f in a \mathbb{C}-algebra 𝒜\mathcal{A} are Murray von Neumann equivalent, written efe\sim f, if there are v,wSv,w\in S with vw=evw=e and wv=fwv=f. Let 𝒜\mathcal{A} be a unital \mathbb{C}-algebra, and let V(𝒜)\mathrm{V}(\mathcal{A}) denote the collection of all Murray-von Neumann equivalence classes of idempotents in M(𝒜)\mathrm{M}_{\infty}(\mathcal{A}). Equip V(𝒜)\mathrm{V}(\mathcal{A}) with the addition operation defined by

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

Then V(𝒜)\mathrm{V}(\mathcal{A}) becomes a commutative monoid. Let 𝒜\mathcal{A} be a unital complex algebra. Then idempotents e,fM(𝒜)e,f\in\mathrm{M}_{\infty}(\mathcal{A}) are Murray-von Neumann equivalent if and only if there is an invertible element uu in some Mn(𝒜)\mathrm{M}_{n}(\mathcal{A}) such that ueu1=fueu^{-1}=f.

Let AA be a unital \mathbb{C}-algebra. The group K0(𝒜)\mathrm{K}_{0}(\mathcal{A}) is the Grothendieck group of V(𝒜)\mathrm{V}(\mathcal{A}). Now, note that K0\mathrm{K}_{0} is functorial for unital C\mathrm{C}^{*}-algebra homomorphisms in a natural way: if ϕ:𝒜B\phi\colon\mathcal{A}\to B is a unital ring homomorphism, and eM(𝒜)e\in\mathrm{M}_{\infty}(\mathcal{A}) is an idempotent, then ϕ(e)\phi(e) is an idempotent in M()\mathrm{M}_{\infty}(\mathcal{B}), and the map [e][ϕ(e)][e]\mapsto[\phi(e)] is well-defined as a map V(𝒜)V()\mathrm{V}(\mathcal{A})\to\mathrm{V}(\mathcal{B}). Hence it induces a map ϕ:K0(𝒜)K0()\phi_{*}\colon\mathrm{K}_{0}(\mathcal{A})\to\mathrm{K}_{0}(\mathcal{B}) by the universal properties of this group

Let 𝒜\mathcal{A} be a not-necessarily unital \mathbb{C}-algebra, and let A~\tilde{A} be its unitisation, which is equipped with a canonical unital \mathbb{C}-algebra homomorphism ϕ:𝒜~\phi\colon\tilde{\mathcal{A}}\to\mathbb{C} with kernel 𝒜\mathcal{A}. Then the group K0(𝒜)\mathrm{K}_{0}(\mathcal{A}) is defined by

K0(𝒜)=defKerϕ:K0(𝒜~)K0().\mathrm{K}_{0}(\mathcal{A})\overset{\mathrm{def}}{=}\operatorname{Ker}\phi_{*}\colon\mathrm{K}_{0}(\tilde{\mathcal{A}})\to\mathrm{K}_{0}(\mathbb{C}).

A local Banach algebra is a normed algebra 𝒜\mathcal{A} which is closed under holomorphic functional calculus (i.e. if x𝒜x\in\mathcal{A} and ff is an analytic function on a neighborhood of the spectrum of xx in the completion of 𝒜\mathcal{A}, with f(0)=0f(0)=0 if 𝒜\mathcal{A} is nonunital, then f(x)𝒜f(x)\in\mathcal{A}.) For technical reasons we will also require that all matrix algebras over AA have the same property. If 𝒜\mathcal{A} is a *-algebra, it will be called a local Banach *-algebra; if the norm is a pre-C\mathrm{C}^{*}-norm, 𝒜\mathcal{A} will be called a local C\mathrm{C}^{*}-algebra.

Recall that GL(𝒜)=deflimGLn(𝒜)\mathrm{GL}_{\infty}(\mathcal{A})\overset{\mathrm{def}}{=}\varinjlim\mathrm{GL}_{n}(\mathcal{A}) and that GL(𝒜)0=deflimGLn(𝒜)0\mathrm{GL}_{\infty}(\mathcal{A})_{0}\overset{\mathrm{def}}{=}\varinjlim\mathrm{GL}_{n}(\mathcal{A})_{0}, where the group GLn(𝒜)0\mathrm{GL}_{n}(\mathcal{A})_{0} is the connected component of the identity. The group K1(𝒜)\mathrm{K}_{1}(\mathcal{A}) is defined by K1(𝒜)=defGL(𝒜)/GL(𝒜)0\mathrm{K}_{1}(\mathcal{A})\overset{\mathrm{def}}{=}\mathrm{GL}_{\infty}(\mathcal{A})/\mathrm{GL}_{\infty}(\mathcal{A})_{0}. It is known [Bl98, p. 59] that the group K1(𝒜)\mathrm{K}_{1}(\mathcal{A}) is countable if the algebra 𝒜\mathcal{A} is separable.

Example 2.1

If 𝒜=C(X)\mathcal{A}=\mathrm{C}(X) is the algebra of a compact Hausdorff space XX, then by [GVF01, Corollary 3.21 p. 101] we have an isomorphism K0(C(X))=K0(X)\mathrm{K}_{0}(\mathrm{C}(X))=\mathrm{K}^{0}(X) where K0(X)\mathrm{K}^{0}(X) is the topological K\mathrm{K}-theory group defined by vector bundles.

Fredholm operators

Following [AbA02, Definition 4.37 p. 156], we say that a bounded linear operator T:XYT\colon X\to Y, acting between complex Banach spaces XX and YY, is a Fredholm operator if the subspaces KerT\operatorname{Ker}T and Y/RanTY/\operatorname{Ran}T are finite-dimensional. In this case, we introduce the index

(2.1) IndexT=defdimKerTdimY/RanT.\operatorname{Index}T\overset{\mathrm{def}}{=}\dim\operatorname{Ker}T-\dim Y/\operatorname{Ran}T.

Every Fredholm operator has a closed range by [AbA02, Lemma 4.38 p. 156]. Recall the Banach version of Atkinson’s theorem [AbA02, Theorem 4.46 p. 161].

Theorem 2.2

A bounded operator T:XYT\colon X\to Y between Banach spaces is a Fredholm operator if and only if there exists a bounded operator R:XYR\colon X\to Y such that RTIdXRT-\mathrm{Id}_{X} and TRIdYTR-\mathrm{Id}_{Y} are compact operators. Moreover, we can replace ¡¡compact¿¿ by ¡¡finite-rank¿¿ in this assertion.

Note that [AbA02, Theorem 4.48 p. 163] the set Fred(X,Y)\mathrm{Fred}(X,Y) of all Fredholm operators from XX into YY is an open subset of B(X,Y)\mathrm{B}(X,Y) and the index function Index:Fred(X,Y)\operatorname{Index}\colon\mathrm{Fred}(X,Y)\to\mathbb{Z} is continuous (hence locally constant). By [AbA02, Corollaey 4.47 p. 162], if T:XYT\colon X\to Y is a Fredholm operator and if K:XYK\colon X\to Y is a compact operator, then A+KA+K is a Fredholm operator and

(2.2) IndA=Ind(A+K)\operatorname{Ind}A=\operatorname{Ind}(A+K)

(invariance under compact perturbations). If T:XYT\colon X\to Y and S:YZS\colon Y\to Z are Fredholm operators then STST is also a Fredholm operator and IndexST=IndexS+IndexT\operatorname{Index}ST=\operatorname{Index}S+\operatorname{Index}T, see [AbA02, Theorem 4.43 p. 158]. By [AbA02, Theorem 4.42 p. 157], a bounded operator is a Fredholm operator if and only if its adjoint is also a Fredholm operator. In this case, we have IndexT=IndexT\operatorname{Index}T^{*}=-\operatorname{Index}T.

Recall that by [AbA02, Theorem 4.54 p. 167] any Fredholm operator T:XYT\colon X\to Y admits a generalized inverse (or pseudo-inverse), i.e. an operator U:YXU\colon Y\to X such that

(2.3) TUT=T.TUT=T.

By the way, the same result says that every generalized inverse UU of TT is also a Fredholm operator and satisfies IndexU=IndexT\operatorname{Index}U=-\operatorname{Index}T. In this case, the proof of [AbA02, Theorem 4.32 p. 166] shows that the operators TU:YYTU\colon Y\to Y and IdUT:XX\mathrm{Id}-UT\colon X\to X are bounded projections on the subspaces RanT\operatorname{Ran}T and KerT\operatorname{Ker}T.

Unbounded operators

The following result is [HvNVW16, Theorem 1.2.4 p. 15].

Proposition 2.3

Let f:ΩXf\colon\Omega\to X be a Bochner integrable function and let TT be a closed linear operator with domain domT\operatorname{dom}T in XX and with values in a Banach space YY. Suppose that ff takes its values in domT\operatorname{dom}T almost everywhere and the almost everywhere defined function Tf:ΩYTf\colon\Omega\to Y is Bochner integrable. Then ff is Bochner integrable as a dom(T)\operatorname{dom}(T)-valued function, Ωfdμ\int_{\Omega}f\mathop{}\mathopen{}\mathrm{d}\mu belongs to dom(T)\operatorname{dom}(T) and

(2.4) T(Ωfdμ)=ΩTfdμ.T\bigg{(}\int_{\Omega}f\mathop{}\mathopen{}\mathrm{d}\mu\bigg{)}=\int_{\Omega}T\circ f\mathop{}\mathopen{}\mathrm{d}\mu.

Bisectorial operators

We refer to [Ege15] and to the books [HvNVW18] and [HvNVW23] for more information on bisectorial operators. For any angle ω(0,π2)\omega\in(0,\frac{\pi}{2}), we consider the open bisector Σωbi=defΣω(Σω)\Sigma_{\omega}^{\mathrm{bi}}\overset{\mathrm{def}}{=}\Sigma_{\omega}\cup(-\Sigma_{\omega}) where Σω=def{z\{0}:|argz|<ω}\Sigma_{\omega}\overset{\mathrm{def}}{=}\big{\{}z\in\mathbb{C}\backslash\{0\}:\>|\arg z|<\omega\big{\}} and Σ0bi=(,)\Sigma_{0}^{\mathrm{bi}}=(-\infty,\infty). Following [HvNVW18, Definition 10.6.1 p. 447] , we say that a closed densely defined operator DD on a Banach space XX is bisectorial of type ω(0,π2)\omega\in(0,\frac{\pi}{2}) if its spectrum σ(D)\sigma(D) is a subset of the closed bisector Σωbi¯\overline{\Sigma^{\mathrm{bi}}_{\omega}} and if the subset {zR(z,D):zΣωbi¯}\big{\{}zR(z,D):z\not\in\overline{\Sigma_{\omega}^{\mathrm{bi}}}\big{\}} is bounded in B(X)\mathrm{B}(X). See Figure 1. The infimum of all ω(0,π2)\omega\in(0,\frac{\pi}{2}) such that DD is bisectorial is called the angle of bisectoriality of DD. The definition of an RR-bisectorial operator is obtained by replacing ¡¡bounded¿¿ by ¡¡RR-bounded¿¿.

ω\omegaσ(D)\sigma(D)

Figure 1: the spectrum of a bisectorial operator DD

By [HvNVW18, p. 447], the operator DD is bisectorial if and only if

(2.5) iρ(D)andsupt+tR(it,D)XX<.\mathrm{i}\mathbb{R}^{*}\subset\rho(D)\quad\text{and}\quad\sup_{t\in\mathbb{R}_{*}^{+}}\left\|tR(\mathrm{i}t,D)\right\|_{X\to X}<\infty.

Similarly, the operator DD is RR-bisectorial if and only if

(2.6) iρ(D)and if the set{tR(it,D):t+} is R-bounded.\mathrm{i}\mathbb{R}^{*}\subset\rho(D)\quad\text{and if the set}\quad\{tR(\mathrm{i}t,D):t\in\mathbb{R}_{+}^{*}\}\text{ is $R$-bounded}.
Example 2.4

If the operator iD\mathrm{i}D generates a strongly continuous group of operators on a Banach space XX then [HvNVW18, Example 10.6.3 p. 448] the operator DD is bisectorial of angle 0. Combined with Stone’s theorem [HvNVW18, Theorem G.6.2 p. 544], we see in particular that unbounded selfadjoint operators on Hilbert spaces are bisectorial of angle 0.

If the operator DD is bisectorial of type ω\omega then by [HvNVW18, Proposition 10.6.2 (2) p. 448] its square D2D^{2} is sectorial of type 2ω2\omega and we have

(2.7) RanD2¯=RanD¯andKerD2=KerD.\overline{\operatorname{Ran}D^{2}}=\overline{\operatorname{Ran}D}\quad\text{and}\quad\operatorname{Ker}D^{2}=\operatorname{Ker}D.

Functional calculus

Consider a bisectorial operator DD on a Banach space XX of type ω(0,π2)\omega\in(0,\frac{\pi}{2}). For any θ(ω,π2)\theta\in(\omega,\frac{\pi}{2}) and any function ff in the space

H0(Σθbi)=def{fH(Σθbi):C,s>0zΣθbi:|f(z)|Cmin{|z|s,|z|s}},\mathrm{H}^{\infty}_{0}(\Sigma_{\theta}^{\mathrm{bi}})\overset{\mathrm{def}}{=}\left\{f\in\mathrm{H}^{\infty}(\Sigma_{\theta}^{\mathrm{bi}}):\>\exists C,s>0\>\forall\>z\in\Sigma_{\theta}^{\mathrm{bi}}:\>|f(z)|\leqslant C\min\{|z|^{s},|z|^{-s}\}\right\},

we can define an operator f(D)f(D) acting on the space XX by integrating on the boundary of a bisector

(2.8) f(D)=def12πiΣνbif(z)R(z,D)dz,f(D)\overset{\mathrm{def}}{=}\frac{1}{2\pi\mathrm{i}}\int_{\partial\Sigma^{\mathrm{bi}}_{\nu}}f(z)R(z,D)\mathop{}\mathopen{}\mathrm{d}z,

using the resolvent operator R(z,D)R(z,D) where ω<ν<θ\omega<\nu<\theta. See [Ege15, Section 3.2.1] for a more precise explanation. The operator DD is said to have a bounded H(Σθbi)\mathrm{H}^{\infty}(\Sigma_{\theta}^{\mathrm{bi}}) functional calculus, if there exists a constant C0C\geqslant 0 such that

(2.9) f(D)XXCfH(Σθbi)\big{\|}f(D)\big{\|}_{X\to X}\leqslant C\left\|f\right\|_{\mathrm{H}^{\infty}(\Sigma_{\theta}^{\mathrm{bi}})}

for any function ff in the space H0(Σθbi)\mathrm{H}^{\infty}_{0}(\Sigma_{\theta}^{\mathrm{bi}}).

Example 2.5

By [HvNVW18, Theorem 10.7.10 p. 461], if the operator iD\mathrm{i}D is the generator of a bounded strongly continuous group on a UMD\mathrm{UMD} Banach space XX then the operator DD admits a H(Σθbi)\mathrm{H}^{\infty}(\Sigma_{\theta}^{\mathrm{bi}}) functional calculus for any angle θ>0\theta>0.

RR-boundedness

Suppose that 1<p<1<p<\infty. Following [HvNVW18, Definition 8.1.1, Remark 8.1.2 p. 165], we say that a set \mathcal{F} of bounded operators on a Banach space XX is RR-bounded provided that there exists a constant C0C\geqslant 0 such that for any finite families T1,,TnT_{1},\ldots,T_{n} in \mathcal{F} and x1,,xnx_{1},\ldots,x_{n} in XX, we have

(2.10) k=1nεkTk(xk)Lp(Ω,X)Ck=1nεkxkLp(Ω,X),\Bigg{\|}\sum_{k=1}^{n}\varepsilon_{k}\otimes T_{k}(x_{k})\Bigg{\|}_{\mathrm{L}^{p}(\Omega,X)}\leqslant C\Bigg{\|}\sum_{k=1}^{n}\varepsilon_{k}\otimes x_{k}\Bigg{\|}_{\mathrm{L}^{p}(\Omega,X)},

where (εk)k1(\varepsilon_{k})_{k\geqslant 1} is a sequence of independent Rademacher variables on some probability space Ω\Omega. This property is independent of pp.

3 Banach Fredholm modules and Banach K\mathrm{K}-homology

In this section, we define a Banach space variant of the theory of K\mathrm{K}-homology, relying on the notion of Fredholm module. We want replace Hilbert spaces by Banach spaces. Note that the classical notion of Fredholm module admit different variations in the literature (compare the references [HiR00, Definition 8.1.1 p. 199], [Con94, Definition 1 p. 293] and [CGIS14, Definition 2.2]). We start to introduce the following definition.

Definition 3.1

Let 𝒜\mathcal{A} be an algebra. An odd Banach Fredholm module (X,F)(X,F) over 𝒜\mathcal{A} on XX consists of a Banach space XX endowed with a representation π:𝒜B(X)\pi\colon\mathcal{A}\to\mathrm{B}(X), and a bounded operator F:XXF\colon X\to X such that

  1. 1.

    F2IdXF^{2}-\mathrm{Id}_{X} is a compact operator on XX,

  2. 2.

    for any a𝒜a\in\mathcal{A} the commutator [F,π(a)][F,\pi(a)] is a compact operator on XX.

Sometimes, we will use the notation [F,a][F,a] for [F,π(a)][F,\pi(a)]. We also introduce a natural notion of even Fredholm module.

Definition 3.2

Let 𝒜\mathcal{A} be an algebra. An even Banach Fredholm module (X,F,γ)(X,F,\gamma) over 𝒜\mathcal{A} on XX consists of a Fredholm module (X,F)(X,F) endowed with a bounded operator γ:XX\gamma\colon X\to X with γ2=IdX\gamma^{2}=\mathrm{Id}_{X} and such that

(3.1) Fγ=γFand[π(a),γ]=0F\gamma=-\gamma F\quad\text{and}\quad[\pi(a),\gamma]=0

for any a𝒜a\in\mathcal{A}. We say that γ\gamma is the grading operator.

In this case, P=defIdX+γ2P\overset{\mathrm{def}}{=}\frac{\mathrm{Id}_{X}+\gamma}{2} is a bounded projection and we can write X=X+XX=X_{+}\oplus X_{-} where X+=defRanPX_{+}\overset{\mathrm{def}}{=}\operatorname{Ran}P and X=defRan(IdXP)X_{-}\overset{\mathrm{def}}{=}\operatorname{Ran}(\mathrm{Id}_{X}-P). With respect to this decomposition, the equations of (3.1) implies that we can write

(3.2) π(a)=[π+(a)00π(a)]andF=[0FF+0],\pi(a)=\begin{bmatrix}\pi_{+}(a)&0\\ 0&\pi_{-}(a)\\ \end{bmatrix}\quad\text{and}\quad F=\begin{bmatrix}0&F_{-}\\ F_{+}&0\\ \end{bmatrix},

where π+:𝒜B(X+)\pi_{+}\colon\mathcal{A}\to\mathrm{B}(X_{+}) and π:𝒜B(X)\pi_{-}\colon\mathcal{A}\to\mathrm{B}(X_{-}) are representations of 𝒜\mathcal{A}. In particular, we have

(3.3) [F,π(a)]=[0Faa+F+F+a+aF+0].\left[F,\pi(a)\right]=\begin{bmatrix}0&F_{-}a_{-}-a_{+}F_{+}\\ F_{+}a_{+}-a_{-}F_{+}&0\\ \end{bmatrix}.
Remark 3.3

Note that if XX is a Hilbert space, these definitions are weaker generalizations of the notions of Fredholm modules of the previous references since the assumption of selfadjointness is not required.

Example 3.4 (from odd to even Fredholm Banach modules)

Let (X,F)(X,F) be an odd Banach Fredholm module over 𝒜\mathcal{A}. It is possible to construct an even Banach Fredholm module (X,F,γ)(X^{\prime},F^{\prime},\gamma) by letting X=defXXX^{\prime}\overset{\mathrm{def}}{=}X\oplus X, π=defππ\pi^{\prime}\overset{\mathrm{def}}{=}\pi\oplus\pi, F=def[0FF0]F^{\prime}\overset{\mathrm{def}}{=}\begin{bmatrix}0&F\\ F&0\\ \end{bmatrix} and γ=def[IdX00IdX]\gamma\overset{\mathrm{def}}{=}\begin{bmatrix}-\mathrm{Id}_{X}&0\\ 0&\mathrm{Id}_{X}\\ \end{bmatrix}. Indeed, we have (F)2=[F200F2][IdX00IdX](F^{\prime})^{2}=\begin{bmatrix}F^{2}&0\\ 0&F^{2}\\ \end{bmatrix}\sim\begin{bmatrix}\mathrm{Id}_{X}&0\\ 0&\mathrm{Id}_{X}\\ \end{bmatrix}. Moreover, the commutator

[F,π(a)]=Fπ(a)π(a)F=[0[F,π(a)][F,π(a)]0]\left[F^{\prime},\pi^{\prime}(a)\right]=F^{\prime}\pi^{\prime}(a)-\pi^{\prime}(a)F^{\prime}=\begin{bmatrix}0&[F,\pi(a)]\\ [F,\pi(a)]&0\\ \end{bmatrix}

is compact. Furthermore, we observe that γ2=[IdX00IdX]\gamma^{2}=\begin{bmatrix}\mathrm{Id}_{X}&0\\ 0&\mathrm{Id}_{X}\\ \end{bmatrix},

Fγ=[0FF0][IdX00IdX]=[0FF0]=[IdX00IdX][0FF0]=γFF^{\prime}\gamma=\begin{bmatrix}0&F\\ F&0\\ \end{bmatrix}\begin{bmatrix}-\mathrm{Id}_{X}&0\\ 0&\mathrm{Id}_{X}\\ \end{bmatrix}=\begin{bmatrix}0&F\\ -F&0\\ \end{bmatrix}=-\begin{bmatrix}-\mathrm{Id}_{X}&0\\ 0&\mathrm{Id}_{X}\\ \end{bmatrix}\begin{bmatrix}0&F\\ F&0\\ \end{bmatrix}=-\gamma F^{\prime}

and finally for any a𝒜a\in\mathcal{A}

[π(a),γ]=[π(a)00π(a)][IdX00IdX][IdX00IdX][π(a)00π(a)]=0.\left[\pi^{\prime}(a),\gamma\right]=\begin{bmatrix}\pi(a)&0\\ 0&\pi(a)\\ \end{bmatrix}\begin{bmatrix}-\mathrm{Id}_{X}&0\\ 0&\mathrm{Id}_{X}\\ \end{bmatrix}-\begin{bmatrix}-\mathrm{Id}_{X}&0\\ 0&\mathrm{Id}_{X}\\ \end{bmatrix}\begin{bmatrix}\pi(a)&0\\ 0&\pi(a)\\ \end{bmatrix}=0.

There is a natural notion of direct sum for Banach Fredholm modules. One takes the direct sum 2\oplus_{2} of the Banach spaces, of the representations, and of the operators. We define the zero Banach Fredholm module with the zero Banach space, the zero representation and the zero operator.

The next definition is straightforward variation of [HiR00, Definition 8.2.1 p. 204].

Definition 3.5

Let (X,F)(X,F) be an odd Banach Fredholm module and let U:YXU\colon Y\to X be an isometric isomorphism. Then (U1πU,Y,U1FU)(U^{-1}\pi U,Y,U^{-1}FU) is also a Banach Fredholm module, and we say that it is isometrically equivalent to (X,F)(X,F).

Similarly to [HiR00, Definition 8.2.2 p. 204], we introduce the following notion of equivalence.

Definition 3.6

Suppose that (X,Ft)(X,F_{t}) is a family of odd Banach Fredholm modules parametrized by t[0,1]t\in[0,1], in which the representation and the Banach space remain constant but the operator FtF_{t} varies with tt. If the function [0,1]B(X)[0,1]\to\mathrm{B}(X), tFtt\mapsto F_{t} is norm continuous, then we say that the family defines an operator homotopy between the odd Banach Fredholm modules (X,F0)(X,F_{0}) and (X,F1)(X,F_{1}), and that these two odd Fredholm modules are operator homotopic.

Definition 3.7 (compact perturbation)

Suppose that (X,F)(X,F) and (X,F)(X,F^{\prime}) are Banach Fredholm modules on the same Banach space XX, and that (FF)π(a)(F-F^{\prime})\pi(a) is compact for all a𝒜a\in\mathcal{A}. In this case, we say that (X,F)(X,F) is a compact perturbation of (X,F)(X,F^{\prime}).

Compact perturbation implies operator homotopy since the linear path from FF to FF^{\prime} defines an operator homotopy.

Similar definitions for the even case are left to the reader. In the spirit of [HiR00, Definition 8.2.5 p. 205], we introduce the next definition.

Definition 3.8 (odd Banach K\mathrm{K}-homology group)

Let \mathscr{B} be a class of Banach spaces stable under countable sums containing the zero space. The Banach K\mathrm{K}-homology group K1(𝒜,)\mathrm{K}^{1}(\mathcal{A},\mathscr{B}) is the abelian group with one generator for each isometric equivalence class [(X,F)][(X,F)] of Banach Fredholm modules over 𝒜\mathcal{A} on any XX of \mathscr{B} subject only to the relations:

  1. 1.

    if (X,F0)(X,F_{0}) and (X,F1)(X,F_{1}) are operator homotopic Fredholm modules then [(X,F0)]=[(X,F1)][(X,F_{0})]=[(X,F_{1})] in K1(𝒜,)\mathrm{K}^{1}(\mathcal{A},\mathscr{B}),

  2. 2.

    if (X,F0)(X,F_{0}) and (X,F1)(X,F_{1}) are any two Fredholm modules then

    [(X,F0)(X,F1)]=[(X,F0)]+[(X,F1)][(X,F_{0})\oplus(X,F_{1})]=[(X,F_{0})]+[(X,F_{1})]

    in K1(𝒜,)\mathrm{K}^{1}(\mathcal{A},\mathscr{B}).

Similarly, we introduce a group K0(𝒜,)\mathrm{K}^{0}(\mathcal{A},\mathscr{B}) in the even case.

Example 3.9

If \mathscr{B} is the class of Hilbert spaces, the previous groups seems different from the K-homology groups K0(𝒜)\mathrm{K}^{0}(\mathcal{A}) and K1(𝒜)\mathrm{K}^{1}(\mathcal{A}) of [HiR00, Definition 8.2.5 p. 205] (defined for a separable C\mathrm{C}^{*}-algebra 𝒜\mathcal{A}) since the bounded operator FF of a Banach Fredholm module (H,F)(H,F) is not necessarily selfadjoint.

Example 3.10

For any 1p<1\leqslant p<\infty, it is natural to consider the class p\mathscr{L}^{p} of Lp\mathrm{L}^{p}-spaces and the class ncp\mathscr{L}^{p}_{\mathrm{nc}} of noncommutative Lp\mathrm{L}^{p}-spaces. We could also consider the class 𝒮p\mathscr{S}\mathscr{L}_{p} of subspaces of Lp\mathrm{L}^{p}-spaces and the class 𝒮𝒬p\mathscr{S}\mathscr{Q}\mathscr{L}_{p} of subspaces of quotients of Lp\mathrm{L}^{p}-spaces.

Definition 3.11 (degenerated Fredholm module)

Let 𝒜\mathcal{A} be an algebra. A odd Fredholm module (X,F)(X,F) over 𝒜\mathcal{A} on XX is said to be degenerate if for any a𝒜a\in\mathcal{A} we have [F,π(a)]=0[F,\pi(a)]=0.

The interest is the following result.

Proposition 3.12

The class defined by a degenerate odd Banach Fredholm module (X,F)(X,F) in the group K1(𝒜,)\mathrm{K}^{1}(\mathcal{A},\mathscr{B}) is zero.

Proof : We introduce (Y,F)(Y,F^{\prime}) where YY is the direct sum i=1X\oplus_{i=1}^{\infty}X of infinitely many copies of XX, and where FF^{\prime} and π\pi^{\prime} are infinite direct sums of copies of FF and π\pi. For any a𝒜a\in\mathcal{A}, we have [F,π(a)]=0[F^{\prime},\pi(a)]=0. Consequently (Y,F)(Y,F^{\prime}) is a Banach Fredholm module. But clearly (X,F)(Y,F)(X,F)\oplus(Y,F^{\prime}) is isometrically equivalent to (Y,F)(Y,F^{\prime}). Hence we have [(X,F)]+[(Y,F)]=[(Y,F)][(X,F)]+[(Y,F^{\prime})]=[(Y,F^{\prime})] in K-homology. We conclude that [(X,F)]=0[(X,F)]=0.    

Proposition 3.13 (even to odd Fredholm module)

Let 𝒜\mathcal{A} be an algebra. Consider an even Banach Fredholm module (X,F,γ)(X,F,\gamma). Let V:XX+V\colon X_{-}\to X_{+} be a surjective isomorphism such that

(3.4) VF+V=F,Vπ(a)=π+(a)V,a𝒜.VF_{+}V=F_{-},\quad V\pi_{-}(a)=\pi_{+}(a)V,\quad a\in\mathcal{A}.

With the notations of (3.2) and G=defVF+:X+X+G\overset{\mathrm{def}}{=}VF_{+}\colon X_{+}\to X_{+}, the pair (X+,G)(X_{+},G) is an odd Banach Fredholm module on 𝒜\mathcal{A}, where the Banach space X+X_{+} is endowed with the homomorphism π+:𝒜B(X+)\pi_{+}\colon\mathcal{A}\to\mathrm{B}(X_{+}).

Proof : Since F2=IdXF^{2}=\mathrm{Id}_{X}, we have FF+=IdX+F_{-}F_{+}=\mathrm{Id}_{X_{+}} and F+F=IdXF_{+}F_{-}=\mathrm{Id}_{X_{-}}. We obtain that the operator

G2IdX+=VF+VF+IdX+=(3.4)FF+IdX+G^{2}-\mathrm{Id}_{X_{+}}=VF_{+}VF_{+}-\mathrm{Id}_{X_{+}}\overset{\eqref{div-986}}{=}F_{-}F_{+}-\mathrm{Id}_{X_{+}}

is compact. Moreover, we have

[VF+,π+(a)]=VF+π+(a)π+(a)VF+=(3.4)VF+π+(a)VV1π+(a)FV1\displaystyle\big{[}VF_{+},\pi_{+}(a)\big{]}=VF_{+}\pi_{+}(a)-\pi_{+}(a)VF_{+}\overset{\eqref{div-986}}{=}VF_{+}\pi_{+}(a)VV^{-1}-\pi_{+}(a)F_{-}V^{-1}
=(3.4)VF+Vπ(a)V1π+(a)FV1=(3.4)Fπ(a)V1π+(a)FV1\displaystyle\overset{\eqref{div-986}}{=}VF_{+}V\pi_{-}(a)V^{-1}-\pi_{+}(a)F_{-}V^{-1}\overset{\eqref{div-986}}{=}F_{-}\pi_{-}(a)V^{-1}-\pi_{+}(a)F_{-}V^{-1}
=(Fπ(a)π+(a)F)V1,\displaystyle=(F_{-}\pi_{-}(a)-\pi_{+}(a)F_{-})V^{-1},

which is a compact operator by (3.3) and since the space of compact operators is an ideal.    

For the proof of Theorem 6.6, we introduce the notation

(3.5) dx=def[F,x]\mathop{}\mathopen{}\mathrm{d}x\overset{\mathrm{def}}{=}[F,x]

(Note that dx=defi[F,x]\mathop{}\mathopen{}\mathrm{d}x\overset{\mathrm{def}}{=}\mathrm{i}[F,x] is better for some topics but we does not need it in this paper). A simple computation shows that

(3.6) d(ab)=adb+dabandFda=daF.\mathop{}\mathopen{}\mathrm{d}(ab)=a\mathop{}\mathopen{}\mathrm{d}b+\mathop{}\mathopen{}\mathrm{d}a\cdot b\quad\text{and}\quad F\mathop{}\mathopen{}\mathrm{d}a=-\mathop{}\mathopen{}\mathrm{d}aF.

We also need a notion of Banach Fredholm module such that the relation F2IdF^{2}-\mathrm{Id} is compact is ¡¡almost satisfied¿¿.

Definition 3.14

Let 𝒜\mathcal{A} be an algebra. A possibly kernel-degenerate Banach Fredholm module (X,F)(X,F) over 𝒜\mathcal{A} on XX consists of a Banach space XX endowed with a representation π:𝒜B(X)\pi\colon\mathcal{A}\to\mathrm{B}(X), and a bounded operator F:XXF\colon X\to X such that

  1. 1.

    XX can be written X=YKerFX=Y\oplus\operatorname{Ker}F for some Banach space YY, i.e. the subspace KerF\operatorname{Ker}F is complemented in XX by a bounded projection Q:XXQ\colon X\to X,

  2. 2.

    F2iYF^{2}-i_{Y} is compact from YY into XX where iY:YXi_{Y}\colon Y\to X is the canonical injection,

  3. 3.

    for any a𝒜a\in\mathcal{A} the commutator [F,π(a)][F,\pi(a)] is a compact operator on YY.

4 Coupling between Banach K-homology with K-theory

We start with the odd case.

Theorem 4.1 (coupling, odd case)

Let 𝒜\mathcal{A} be a unital algebra and let \mathscr{B} be a class of Banach spaces stable under countable sums and containing the zero space. Consider an invertible uu of the matrix algebra Mn(𝒜)\mathrm{M}_{n}(\mathcal{A}) and an odd Banach Fredholm module (X,F)(X,F) over 𝒜\mathcal{A} with π(1)=IdX\pi(1)=\mathrm{Id}_{X}. We introduce the bounded operators Pn=defIdId+F2:n2(X)n2(X)P_{n}\overset{\mathrm{def}}{=}\mathrm{Id}\otimes\frac{\mathrm{Id}+F}{2}\colon\ell^{2}_{n}(X)\to\ell^{2}_{n}(X) and un=def(Idπ)(u):n2(X)n2(X)u_{n}\overset{\mathrm{def}}{=}(\mathrm{Id}\otimes\pi)(u)\colon\ell^{2}_{n}(X)\to\ell^{2}_{n}(X). Then the bounded operator

(4.1) PnunPn(IdPn):n2(X)n2(X)P_{n}u_{n}P_{n}-(\mathrm{Id}-P_{n})\colon\ell^{2}_{n}(X)\to\ell^{2}_{n}(X)

is Fredholm. If XX is in \mathscr{B}, its Fredholm index depends only on [u]K1(𝒜)[u]\in\mathrm{K}_{1}(\mathcal{A}) and on [F]K1(𝒜,)[F]\in\mathrm{K}^{1}(\mathcal{A},\mathscr{B}).

Proof : We only do the case n=1n=1 and we use the shorthand notation PP for the map P1=Id+F2P_{1}=\frac{\mathrm{Id}+F}{2}. The case n>1n>1 is left to the reader. Since

(4.2) P2=Id+2F+F24=Id+2F4+14F2Id+2F4+14Id=Id+F2=P.P^{2}=\frac{\mathrm{Id}+2F+F^{2}}{4}=\frac{\mathrm{Id}+2F}{4}+\frac{1}{4}F^{2}\sim\frac{\mathrm{Id}+2F}{4}+\frac{1}{4}\mathrm{Id}=\frac{\mathrm{Id}+F}{2}=P.

the map PP is a projection modulo a compact operator. If u𝒜u\in\mathcal{A} is invertible, we have

[Pπ(u)P(IdP)][Pπ(u1)P(IdP)]\displaystyle\big{[}P\pi(u)P-(\mathrm{Id}-P)\big{]}\big{[}P\pi(u^{-1})P-(\mathrm{Id}-P)\big{]}
=Pπ(u)P2Pπ(u1)PPπ(u)P(IdP)0(IdP)P0π(u1)P+(IdP)2IP\displaystyle=P\pi(u)\underbrace{P^{2}}_{\sim P}\pi(u^{-1})P-P\pi(u)\underbrace{P(\mathrm{Id}-P)}_{\sim 0}-\underbrace{(\mathrm{Id}-P)P}_{\sim 0}\pi(u^{-1})P+\underbrace{(\mathrm{Id}-P)^{2}}_{\mathrm{I}-P}
Pπ(u)Pπ(u1)P+(IdP)=Pπ(u)[Pπ(u1)π(u1)P+π(u1)P]P+(IdP)\displaystyle\sim P\pi(u)P\pi(u^{-1})P+(\mathrm{Id}-P)=P\pi(u)\big{[}P\pi(u^{-1})-\pi(u^{-1})P+\pi(u^{-1})P\big{]}P+(\mathrm{Id}-P)
=Pπ(u)([P,π(u1)]+π(u1)P)P+(IdP)\displaystyle=P\pi(u)\big{(}[P,\pi(u^{-1})]+\pi(u^{-1})P\big{)}P+(\mathrm{Id}-P)
=Pπ(u)(12[F,π(u1)]+π(u1)P)P+(IdP)\displaystyle=P\pi(u)\big{(}\tfrac{1}{2}[F,\pi(u^{-1})]+\pi(u^{-1})P\big{)}P+(\mathrm{Id}-P)
(4.2)12Pπ(u)[F,π(u1)]P+Pπ(uu1)P+(IdP)\displaystyle\overset{\eqref{P2=P}}{\sim}\tfrac{1}{2}P\pi(u)[F,\pi(u^{-1})]P+P\pi(uu^{-1})P+(\mathrm{Id}-P)
=12Pπ(u)[F,π(u1)]compactP+P2+(IdP)Id.\displaystyle=\tfrac{1}{2}P\pi(u)\underbrace{[F,\pi(u^{-1})]}_{\textrm{compact}}P+P^{2}+(\mathrm{Id}-P)\sim\mathrm{Id}.

Similarly, we have [Pπ(u1)P(IdP)][Pπ(u)P(IdP)]Id\big{[}P\pi(u^{-1})P-(\mathrm{Id}-P)\big{]}\big{[}P\pi(u)P-(\mathrm{Id}-P)\big{]}\sim\mathrm{Id}. By Atkinson’s theorem (Theorem 2.2), we deduce that the bounded operator Pπ(u)P(IdP)P\pi(u)P-(\mathrm{Id}-P) is Fredholm. The other verifications are routine.   

So we have a pairing ,K1(𝒜),K1(𝒜,):K1(𝒜)×K1(𝒜,)\langle\cdot,\cdot\rangle_{\mathrm{K}_{1}(\mathcal{A}),\mathrm{K}^{1}(\mathcal{A},\mathscr{B})}\colon\mathrm{K}_{1}(\mathcal{A})\times\mathrm{K}^{1}(\mathcal{A},\mathscr{B})\to\mathbb{Z} defined by

(4.3) [u],[(X,F)]K1(𝒜),K1(𝒜,)=IndexPnunPn(IdPn)=IndexPuP(IdP).\big{\langle}[u],[(X,F)]\big{\rangle}_{\mathrm{K}_{1}(\mathcal{A}),\mathrm{K}^{1}(\mathcal{A},\mathscr{B})}=\operatorname{Index}P_{n}u_{n}P_{n}-(\mathrm{Id}-P_{n})=\operatorname{Index}PuP-(\mathrm{Id}-P).

If P2=PP^{2}=P, a similar proof shows that

(4.4) [u],[(X,F)]K1(𝒜),K1(𝒜,)=IndexPnunPn:Pn(n2(X)))Pn(n2(X)).\big{\langle}[u],[(X,F)]\big{\rangle}_{\mathrm{K}_{1}(\mathcal{A}),\mathrm{K}^{1}(\mathcal{A},\mathscr{B})}=\operatorname{Index}P_{n}u_{n}P_{n}\colon P_{n}(\ell^{2}_{n}(X)))\to P_{n}(\ell^{2}_{n}(X)).

Similarly, we prove the even case.

Theorem 4.2 (coupling, even case)

Let 𝒜\mathcal{A} be a unital algebra and let \mathscr{B} be a class of Banach spaces stable under countable sums and containing the zero space. Consider a projection ee of the matrix algebra Mn(𝒜)\mathrm{M}_{n}(\mathcal{A}) and an even Banach Fredholm module (X+X,[0FF+0],γ)\bigg{(}X_{+}\oplus X_{-},\begin{bmatrix}0&F_{-}\\ F_{+}&0\\ \end{bmatrix},\gamma\bigg{)} over 𝒜\mathcal{A} with π(1)=IdX\pi(1)=\mathrm{Id}_{X}. We introduce the bounded operator en=def(Idπ)(e):n2(X)n2(X)e_{n}\overset{\mathrm{def}}{=}(\mathrm{Id}\otimes\pi)(e)\colon\ell^{2}_{n}(X)\to\ell^{2}_{n}(X). Then the bounded operator

(4.5) en(IdF+)en:en(n2(X+))en(n2(X))e_{n}(\mathrm{Id}\otimes F_{+})e_{n}\colon e_{n}(\ell^{2}_{n}(X_{+}))\to e_{n}(\ell^{2}_{n}(X_{-}))

is Fredholm. If XX is in \mathscr{B}, its Fredholm index depends only on [e]K0(𝒜)[e]\in\mathrm{K}_{0}(\mathcal{A}) and on [F]K0(𝒜,)[F]\in\mathrm{K}^{0}(\mathcal{A},\mathscr{B}).

Proof : We only do the case n=1n=1 and we use the shorthand notation e=e1e=e_{1}. The case n>1n>1 is left to the reader. Note that the operators FF+IdX+F_{-}F_{+}-\mathrm{Id}_{X_{+}} and F+FIdXF_{+}F_{-}-\mathrm{Id}_{X_{-}} are compact since [FF+00F+F]IdX=F2IdX\begin{bmatrix}F_{-}F_{+}&0\\ 0&F_{+}F_{-}\\ \end{bmatrix}-\mathrm{Id}_{X}=F^{2}-\mathrm{Id}_{X} is compact. We have

(eF+e)(eFe)=eF+eFe=e(F+eeF++eF+)Fe\displaystyle(eF_{+}e)(eF_{-}e)=eF_{+}eF_{-}e=e(F_{+}e-eF_{+}+eF_{+})F_{-}e
=e([F+,e]+eF+)Fe=e[F+,e]compactFe+eF+Fee.\displaystyle=e([F_{+},e]+eF_{+})F_{-}e=e\underbrace{[F_{+},e]}_{\textrm{compact}}F_{-}e+eF_{+}F_{-}e\sim e.

Similarly, we have (eFe)(eF+e)e(eF_{-}e)(eF_{+}e)\sim e. By Atkinson’s theorem (Theorem 2.2), we deduce that the bounded operator eF+e:e(X+)en(X)eF_{+}e\colon e(X_{+})\to e_{n}(X_{-}) is Fredholm. The proof is complete.    

Consequently, we have a pairing ,K0(𝒜),K0(𝒜,):K0(𝒜)×K0(𝒜,)\langle\cdot,\cdot\rangle_{\mathrm{K}_{0}(\mathcal{A}),\mathrm{K}^{0}(\mathcal{A},\mathscr{B})}\colon\mathrm{K}_{0}(\mathcal{A})\times\mathrm{K}^{0}(\mathcal{A},\mathscr{B})\to\mathbb{Z} defined by

(4.6) [e],[(X,F,γ)]K0(𝒜),K0(𝒜,)=Indexen(IdF+)en.\big{\langle}[e],[(X,F,\gamma)]\big{\rangle}_{\mathrm{K}_{0}(\mathcal{A}),\mathrm{K}^{0}(\mathcal{A},\mathscr{B})}=\operatorname{Index}e_{n}(\mathrm{Id}\otimes F_{+})e_{n}.
Remark 4.3

Similarly to [Con94, Proposition 2 p. 289], it is easy to extend these result to the non-unital case. For example, in the odd case, it suffices to replace Mn(𝒜)\mathrm{M}_{n}(\mathcal{A}) by Mn(𝒜~)\mathrm{M}_{n}(\tilde{\mathcal{A}}) where 𝒜~\tilde{\mathcal{A}} is the unitization of the non-unital algebra 𝒜\mathcal{A}.

5 Summability of Banach Fredholm modules

Approximation numbers

Following [Kon86, Definition p. 68], we say that a map assigning to any bounded operator TT between Banach spaces a sequence ((sn(T))((s_{n}(T)) of real numbers is an ss-number function if the following conditions are satisfied.

  1. 1.

    Ts1(T)s2(T)0\left\|T\right\|\geqslant s_{1}(T)\geqslant s_{2}(T)\geqslant\cdots\geqslant 0.

  2. 2.

    If R,T:XYR,T\colon X\to Y and if n,m1n,m\geqslant 1 then sn+m1(R+T)sn(R)+sn(T)s_{n+m-1}(R+T)\leqslant s_{n}(R)+s_{n}(T).

  3. 3.

    If T:X0XT\colon X_{0}\to X, S:XYS\colon X\to Y and R:YY0R\colon Y\to Y_{0} we have

    (5.1) sn(RST)Rsn(S)T.s_{n}(RST)\leqslant\left\|R\right\|s_{n}(S)\left\|T\right\|.
  4. 4.

    We have sn(T)=0s_{n}(T)=0 if rankT<n\operatorname{rank}T<n and sn(Idn2)=1s_{n}(\mathrm{Id}_{\ell^{2}_{n}})=1.

This concept was introduced by Pietsch in [Pie74]. If XX and YY are Hilbert spaces and if T:XYT\colon X\to Y is compact then by [Kon01, p. 69] these numbers coincide with the singular numbers of the operator TT. We refer to the books [Pi80], [Pie87], [Kon01] and to the survey papers [Kon01] and [Pie23] for more information. The history of this topic is equally described in [Pie07, Chapter 6].

Example 5.1

The approximation numbers [Kon86, Definition I.d.14 p. 69] of a bounded linear operator T:XYT\colon X\to Y are defined by

an(T)=definf{TR:R:XY,rankR<n},n1.a_{n}(T)\overset{\mathrm{def}}{=}\inf\{\left\|T-R\right\|:R\colon X\to Y,\operatorname{rank}R<n\},\quad n\geqslant 1.

The Weyl numbers are defined by

(5.2) xn(T)=sup{an(TA):A:2XA=1},n1.x_{n}(T)=\sup\{a_{n}(TA):A\colon\ell^{2}\to X\left\|A\right\|=1\},\quad n\geqslant 1.

These sequences are example of ss-number sequences by [Kon86, Lemma p. 69]. Moreover, we have xn(T)an(T)x_{n}(T)\leqslant a_{n}(T) for any integer n1n\geqslant 1. Actually, the approximation numbers are the largest ss-numbers by [Kon86, Lemma p. 69].

Quasi-Banach ideals

Following [Kon86, Definition I.d.1 p. 56], a quasi-Banach ideal of operators (𝔄,α)(\mathfrak{A},\alpha) is a subclass 𝔄\mathfrak{A} of all bounded linear operators between Banach spaces together with α:𝔄+\alpha\colon\mathfrak{A}\to\mathbb{R}^{+} such that for all Banach spaces X,YX,Y the sets 𝔄(X,Y)=def𝔄B(X,Y)\mathfrak{A}(X,Y)\overset{\mathrm{def}}{=}\mathfrak{A}\cap\mathrm{B}(X,Y) satisfy:

  1. 1.

    𝔄(X,Y)\mathfrak{A}(X,Y) contains all finite-rank operators from XX to YY and α(Id)=1\alpha(\mathrm{Id}_{\mathbb{C}})=1.

  2. 2.

    (𝔄(X,Y),α)(\mathfrak{A}(X,Y),\alpha) is a quasi-Banach space with quasi-triangle constant KK independent of XX and YY, i.e.

    α(R+T)K(α(R)+α(T)),R,T𝔄(X,Y)\alpha(R+T)\leqslant K(\alpha(R)+\alpha(T)),\quad R,T\in\mathfrak{A}(X,Y)
  3. 3.

    If RB(X0,X)R\in\mathrm{B}(X_{0},X), T𝔄(X,Y)T\in\mathfrak{A}(X,Y), SB(Y,Y0)S\in\mathrm{B}(Y,Y_{0}) for some Banach spaces X0X_{0}, Y0Y_{0}, then

    (5.3) STR𝔄(X0,Y0)andα(STR)Sα(T)R.STR\in\mathfrak{A}(X_{0},Y_{0})\quad\text{and}\quad\alpha(STR)\leqslant\left\|S\right\|\alpha(T)\left\|R\right\|.

If α\alpha is a norm on each 𝔄(X,Y)\mathfrak{A}(X,Y), i.e. K=1K=1, then (𝔄,α)(\mathfrak{A},\alpha) is called Banach ideal of operators, α\alpha the ideal norm. We also define 𝔄(X)=def𝔄(X,X)\mathfrak{A}(X)\overset{\mathrm{def}}{=}\mathfrak{A}(X,X).

Quasi-Banach ideals associated to ss-numbers

Suppose that 0<q<0<q<\infty. Let T(sn(T))n0T\mapsto(s_{n}(T))_{n\geqslant 0} be an ss-number sequence. For any Banach spaces XX and YY, following [Kon86, Definition 1.d.18 p. 72] we define the class

(5.4) Ssq(X,Y)=def{T:XY:(sn(T))q}.S^{q}_{s}(X,Y)\overset{\mathrm{def}}{=}\big{\{}T\colon X\to Y:(s_{n}(T))\in\ell^{q}\big{\}}.

Moreover, if TSsq(X,Y)T\in S^{q}_{s}(X,Y), we let

TSsq(X,Y)=def(sn(T))q.\left\|T\right\|_{S^{q}_{s}(X,Y)}\overset{\mathrm{def}}{=}\left\|(s_{n}(T))\right\|_{\ell^{q}}.

By [Kon86, Lemma p. 72], (Ssq,Ssq)(S^{q}_{s},\left\|\cdot\right\|_{S^{q}_{s}}) is a quasi-Banach ideal. For the cases sn=ans_{n}=a_{n} or if sn=xns_{n}=x_{n} we will use the notations SappqS^{q}_{\mathrm{app}} and SweylqS^{q}_{\mathrm{weyl}}. If X=YX=Y, we use the notation Ssq(X)S^{q}_{s}(X). If in addition XX is equal to a Hilbert space HH, we recover the Schatten class Sq(H)S^{q}(H). If 0<p<q<0<p<q<\infty, we have Ssp(X)Ssq(X)S^{p}_{s}(X)\subset S^{q}_{s}(X). Furthermore, if 1r=1p+1q\frac{1}{r}=\frac{1}{p}+\frac{1}{q}, we have by [Kon86, Proposition 1.d.19 p. 73]

(5.5) Ssp(X)Ssq(X)Ssr(X).S^{p}_{s}(X)\circ S^{q}_{s}(X)\subset S^{r}_{s}(X).

By [Kon80, p. 220], the trace Tr\operatorname{Tr} on the space of finite-rank operators acting on XX admits a unique continuous extension on Sapp1(X)S^{1}_{\mathrm{app}}(X), again denoted Tr\operatorname{Tr}. The same thing is true for Sweyl1S^{1}_{\mathrm{weyl}} for the class of Banach spaces with the bounded approximation property. Moreover, by [Kon86, pp. 224-225]), the previously trace coincide with the sum of eigenvalues of T, i.e. TrT=n=0λn(T)\operatorname{Tr}T=\sum_{n=0}^{\infty}\lambda_{n}(T). Furthermore, if TSapp1(X)T\in S^{1}_{\mathrm{app}}(X) and if R:XXR\colon X\to X is a bounded operator then by [Kon86, Corollary 2 p. 228]

(5.6) Tr(TR)=Tr(RT).\operatorname{Tr}(TR)=\operatorname{Tr}(RT).

It is worth noting that by [Kon86, Theorem 4.b.12 p. 245] a Banach space XX is isomorphic to a Hilbert space if and only if the space 𝒩(X)\mathcal{N}(X) of nuclear operators coincide with the space Sapp1(X)S^{1}_{\mathrm{app}}(X).

Recall the Weyl’s inequality for operators acting on Banach spaces. For that recall that a bounded operator T:XXT\colon X\to X is a Riesz operator if

  1. 1.

    for all λ{0}\lambda\in\mathbb{C}-\{0\}, TλIdT-\lambda\mathrm{Id} is a Fredholm operator and has finite ascent and finite descent,

  2. 2.

    all non-zero spectral values λσ(T)\lambda\in\sigma(T) are eigenvalues of finite multiplicity and have no accumulation point except possibly zero.

Then by [Kon86, Theorem 2.a.6 p. 85] any bounded operator TSweylq(X)T\in S^{q}_{\mathrm{weyl}}(X) is a Riesz operator such that its sequence (λn)(\lambda_{n}) of eigenvalues belongs to the space q\ell^{q} and we have

(5.7) (λn)q21q2eTSweylq(X).\left\|(\lambda_{n})\right\|_{\ell^{q}}\leqslant 2^{\frac{1}{q}}\sqrt{2\mathrm{e}}\left\|T\right\|_{S^{q}_{\mathrm{weyl}}(X)}.

Summability of Banach Fredholm modules

Here, we give a generalization of [Con94, Definition 3 p. 290].

Definition 5.2

Suppose that 0<q<0<q<\infty. We say an odd Banach Fredholm module (X,F)(X,F) or an even Banach Fredhlom (X,F,γ)(X,F,\gamma) is qq-summable if the commutator [F,π(a)][F,\pi(a)] belongs to the space Sappq(X)S^{q}_{\mathrm{app}}(X) for any a𝒜a\in\mathcal{A}.

By (5.5), this implies that if n+1qn+1\geqslant q every product [F,a0][F,a1][F,an][F,a_{0}][F,a_{1}]\cdots[F,a_{n}] of commutators belongs to the space Sapp1(X)S^{1}_{\mathrm{app}}(X).

6 The Chern character of a qq-summable Banach Fredhlom module

In this section, we extend the notion of Chern character introduced by Connes to our setting. In the spirit of [Con94, Definition 3 p. 295] (see also [CPR11, Definition 4.13 p. 34]), we introduce the following definition.

Definition 6.1 (odd case)

Consider a qq-summable odd Banach Fredholm module (X,F)(X,F) for some q>0q>0 over an algebra 𝒜\mathcal{A}. Let nn be an odd integer with n+1qn+1\geqslant q. We define the Chern character by the formula

(6.1) ChnF(a0,a1,,an)=defcnTr(F[F,a0][F,a1][F,an]),a0,a1,,an𝒜,\mathrm{Ch}_{n}^{F}(a_{0},a_{1},\dots,a_{n})\overset{\mathrm{def}}{=}c_{n}\operatorname{Tr}(F[F,a_{0}][F,a_{1}]\cdots[F,a_{n}]),\qquad a_{0},a_{1},\dots,a_{n}\in\mathcal{A},

where cnc_{n} is a constant (but the exact value is not useful for this paper).

In the even case, the definition is slightly different.

Definition 6.2 (even case)

Consider a qq-summable even Banach Fredholm module (X,F,γ)(X,F,\gamma) for some q>0q>0 over an algebra 𝒜\mathcal{A}. Let nn be an even integer with n+1qn+1\geqslant q. We define the Chern character by the formula

(6.2) ChnF(a0,a1,,an)=defcnTr(γF[F,a0][F,a1][F,an]),a0,a1,,an𝒜,\mathrm{Ch}_{n}^{F}(a_{0},a_{1},\dots,a_{n})\overset{\mathrm{def}}{=}c_{n}\operatorname{Tr}(\gamma F[F,a_{0}][F,a_{1}]\cdots[F,a_{n}]),\qquad a_{0},a_{1},\dots,a_{n}\in\mathcal{A},

where we use a constant cnc_{n} (but the exact value is not useful for this paper).

We need the following Banach space generalization of [GVF01, p. 143].

Proposition 6.3

Let XX be a Banach space. Let T:XXT\colon X\mapsto X be a Fredholm operator and let R:XXR\colon X\mapsto X be a bounded operator such that the operators IdTR\mathrm{Id}-TR and IdRT\mathrm{Id}-RT belong to the space Sapp1(X)S^{1}_{\mathrm{app}}(X). Then

IndT=Tr(IdTR)Tr(IdRT).\operatorname{Ind}T=\operatorname{Tr}(\mathrm{Id}-TR)-\operatorname{Tr}(\mathrm{Id}-RT).

Proof : Let UU be a generalized inverse of TT as in (2.3). Then

R=U+(IdUT)RU(IdTR)=defG.R=U+\underbrace{(\mathrm{Id}-UT)R-U(\mathrm{Id}-TR)}_{\overset{\mathrm{def}}{=}G}.

Since IdUT\mathrm{Id}-UT has finite rank and IdTR\mathrm{Id}-TR belongs to the space Sapp1(X)S^{1}_{\mathrm{app}}(X), it follows that GG also belongs to the space Sapp1(X)S^{1}_{\mathrm{app}}(X). Observe that

RTTR=(U+G)TT(U+G)=UTTU+GTTG.RT-TR=(U+G)T-T(U+G)=UT-TU+GT-TG.

We note that TrGT=(5.6)TrTG\operatorname{Tr}GT\overset{\eqref{trace-prop}}{=}\operatorname{Tr}TG. Consequently, it suffices to prove the result for R=UR=U.

Recall that IdTU\mathrm{Id}-TU and IdUT\mathrm{Id}-UT are projections of finite rank. Thus IdTU\mathrm{Id}-TU and IdUT\mathrm{Id}-UT belong to Sapp1(X)S^{1}_{\mathrm{app}}(X). Hence TUUTTU-UT belongs to Sapp1(X)S^{1}_{\mathrm{app}}(X) and recalling that the trace of a finite rank projection is equal to the rank of the projection, we obtain

Tr(IdUT)Tr(IdTU)=rank(IdUT)rank(IdTU)\displaystyle\operatorname{Tr}(\mathrm{Id}-UT)-\operatorname{Tr}(\mathrm{Id}-TU)=\operatorname{rank}(\mathrm{Id}-UT)-\operatorname{rank}(\mathrm{Id}-TU)
=dimKerTdim(Y/RanT)=IndT.\displaystyle=\dim\operatorname{Ker}T-\dim(Y/\operatorname{Ran}T)=\operatorname{Ind}T.

   

Remark 6.4

Let T:XXT\colon X\to X be a bounded operator. At the time of writing, it is unclear if the existence of a bounded operator R:XXR\colon X\mapsto X be a bounded operator such that the operators IdTR\mathrm{Id}-TR and IdRT\mathrm{Id}-RT belong to the space Sapp1(X)S^{1}_{\mathrm{app}}(X) implies that TT is a Fredholm operator.

The following is a Banach space generalization of [GVF01, Proposition 4.2 p. 144].

Corollary 6.5

Let XX be a Banach space. Let T:XXT\colon X\mapsto X be a Fredholm operator and let R:XXR\colon X\mapsto X be a bounded operator such that the operators (IdTR)n(\mathrm{Id}-TR)^{n} and (IdRT)n(\mathrm{Id}-RT)^{n} belong to the space Sapp1(X)S^{1}_{\mathrm{app}}(X) for some integer n1n\geqslant 1. Then the index of the Fredholm operator TT is given by

(6.3) IndT=Tr(IdRT)nTr(IdTR)n.\operatorname{Ind}T=\operatorname{Tr}(\mathrm{Id}-RT)^{n}-\operatorname{Tr}(\mathrm{Id}-TR)^{n}.

Proof : We let K=defIdRTK\overset{\mathrm{def}}{=}\mathrm{Id}-RT and L=defIdTRL\overset{\mathrm{def}}{=}\mathrm{Id}-TR. Note that TK=TTRT=LTTK=T-TRT=LT. We introduce the sum Sn=j=0n1KjRS_{n}=\sum_{j=0}^{n-1}K^{j}R. We have

SnT=j=0n1KjRT=j=0n1Kj(IdK)=IdKn.S_{n}T=\sum_{j=0}^{n-1}K^{j}RT=\sum_{j=0}^{n-1}K^{j}(\mathrm{Id}-K)=\mathrm{Id}-K^{n}.

Moreover, using TK=LTTK=LT in the second equality, we see that

TSn=Tj=0n1KjR=j=0n1LjTR=j=0n1Lj(IdL)=IdLn.TS_{n}=T\sum_{j=0}^{n-1}K^{j}R=\sum_{j=0}^{n-1}L^{j}TR=\sum_{j=0}^{n-1}L^{j}(\mathrm{Id}-L)=\mathrm{Id}-L^{n}.

Hence the operators IdSnT=Kn=(IdRT)n\mathrm{Id}-S_{n}T=K^{n}=(\mathrm{Id}-RT)^{n} and IdTSn=Ln=(IdTR)n\mathrm{Id}-TS_{n}=L^{n}=(\mathrm{Id}-TR)^{n} belong to the space Sapp1S^{1}_{\mathrm{app}}. Consequently, using Proposition 6.3 in the first equality, we obtain

IndT=Tr(IdSnT)Tr(IdTSn)=TrKnTrLn=Tr(IdRT)nTr(IdTR)n.\operatorname{Ind}T=\operatorname{Tr}(\mathrm{Id}-S_{n}T)-\operatorname{Tr}(\mathrm{Id}-TS_{n})=\operatorname{Tr}K^{n}-\operatorname{Tr}L^{n}=\operatorname{Tr}(\mathrm{Id}-RT)^{n}-\operatorname{Tr}(\mathrm{Id}-TR)^{n}.

   

Now we describe the pairing (4.4) in the odd case.

Theorem 6.6 (odd case)

Let (X,F)(X,F) be a qq-summable odd Banach Fredholm module over the algebra 𝒜\mathcal{A} for some q>0q>0 with F2=IdF^{2}=\mathrm{Id}. Consider a class \mathscr{B} of Banach spaces stable under countable sums and containing the zero space. For any element [u][u] of the group K1(𝒜)\mathrm{K}_{1}(\mathcal{A}) and any odd integer nn with n+1qn+1\geqslant q, we have

(6.4) [u],[(X,F)]K1(𝒜),K1(𝒜,)=(1)n+124n+12cnChnF(u1,u,,u1,u).\big{\langle}[u],[(X,F)]\big{\rangle}_{\mathrm{K}_{1}(\mathcal{A}),\mathrm{K}^{1}(\mathcal{A},\mathscr{B})}=\frac{(-1)^{\frac{n+1}{2}}}{4^{\frac{n+1}{2}}c_{n}}\mathrm{Ch}_{n}^{F}(u^{-1},u,\ldots,u^{-1},u).

Proof : We let N=defn+12N\overset{\mathrm{def}}{=}\frac{n+1}{2} and P=defI+F2P\overset{\mathrm{def}}{=}\frac{\mathrm{I}+F}{2}. Note that

P[P,u1]P=P(Pu1u1P)P=P(Pu1+Pu1PPu1Pu1P)P\displaystyle P[P,u^{-1}]P=P(Pu^{-1}-u^{-1}P)P=P(Pu^{-1}+Pu^{-1}P-Pu^{-1}P-u^{-1}P)P
=P(Pu1+u1PPu1u1P)P=P[P,u1]P+P[P,u1]P.\displaystyle=P(Pu^{-1}+u^{-1}P-Pu^{-1}-u^{-1}P)P=P[P,u^{-1}]P+P[P,u^{-1}]P.

Hence P[P,u1]P=0P[P,u^{-1}]P=0 which is equivalent to Pdu1P=0P\mathop{}\mathopen{}\mathrm{d}u^{-1}P=0. Using this equality, in the last equality, we obtain

(6.5) PPu1PuP=Pu1uPPu1PuP=P(u1PPu1)uP\displaystyle P-Pu^{-1}PuP=Pu^{-1}uP-Pu^{-1}PuP=-P(u^{-1}P-Pu^{-1})uP
=P([P,u1])uP=12Pdu1uP=12Pdu1(uPPu+Pu)\displaystyle=P([P,u^{-1}])uP=\frac{1}{2}P\mathop{}\mathopen{}\mathrm{d}u^{-1}uP=\frac{1}{2}P\mathop{}\mathopen{}\mathrm{d}u^{-1}(uP-Pu+Pu)
=14Pdu1du+12Pdu1Pu=14Pdu1du.\displaystyle=-\frac{1}{4}P\mathop{}\mathopen{}\mathrm{d}u^{-1}\mathop{}\mathopen{}\mathrm{d}u+\frac{1}{2}P\mathop{}\mathopen{}\mathrm{d}u^{-1}Pu=-\frac{1}{4}P\mathop{}\mathopen{}\mathrm{d}u^{-1}\mathop{}\mathopen{}\mathrm{d}u.

Since the the Banach Fredholm module is qq-summable, the elements du\mathop{}\mathopen{}\mathrm{d}u and du1\mathop{}\mathopen{}\mathrm{d}u^{-1} belongs to the space Sappq(X)S^{q}_{\mathrm{app}}(X). We deduce that PPu1PuPP-Pu^{-1}PuP belongs to Sappq2(X)S^{\frac{q}{2}}_{\mathrm{app}}(X), hence to SappN(X)S^{N}_{\mathrm{app}}(X) since n+1qn+1\geqslant q. Similarly, we can prove that

(6.6) PPuPu1P=14Pdudu1.P-PuPu^{-1}P=-\frac{1}{4}P\mathop{}\mathopen{}\mathrm{d}u\mathop{}\mathopen{}\mathrm{d}u^{-1}.

Consequently, this operator also belongs to the space SappN(X)S^{N}_{\mathrm{app}}(X). So we use Corollary 6.5 with the operators PuP:P(X)P(X)PuP\colon P(X)\to P(X) and Pu1P:P(X)P(X)Pu^{-1}P\colon P(X)\to P(X) replacing TT and RR. Using the equalities Tr(du1du)N=Tr(dudu1)N\operatorname{Tr}(\mathop{}\mathopen{}\mathrm{d}u^{-1}\mathop{}\mathopen{}\mathrm{d}u)^{N}=\operatorname{Tr}(\mathop{}\mathopen{}\mathrm{d}u\mathop{}\mathopen{}\mathrm{d}u^{-1})^{N} and TrF(dudu1)N=(3.6)TrF(du1du)N\operatorname{Tr}F(\mathop{}\mathopen{}\mathrm{d}u\mathop{}\mathopen{}\mathrm{d}u^{-1})^{N}\overset{\eqref{prop-quantized-deriv}}{=}-\operatorname{Tr}F(\mathop{}\mathopen{}\mathrm{d}u^{-1}\mathop{}\mathopen{}\mathrm{d}u)^{N}, we obtain

Index(PuP)=(6.3)Tr(PPu1PuP)NTr(PPuPu1P)N\displaystyle\operatorname{Index}(PuP)\overset{\eqref{index-power-N}}{=}\operatorname{Tr}(P-Pu^{-1}PuP)^{N}-\operatorname{Tr}(P-PuPu^{-1}P)^{N}
=(6.5)(6.6)Tr(14Pdu1du)NTr(14Pdudu1)N\displaystyle\overset{\eqref{div96}\eqref{div90}}{=}\operatorname{Tr}(-\tfrac{1}{4}P\mathop{}\mathopen{}\mathrm{d}u^{-1}\mathop{}\mathopen{}\mathrm{d}u)^{N}-\operatorname{Tr}(-\tfrac{1}{4}P\mathop{}\mathopen{}\mathrm{d}u\mathop{}\mathopen{}\mathrm{d}u^{-1})^{N}
=(1)N4N[Tr(Pdu1du)NTr(Pdudu1)N]=(1)N4N[TrP(du1du)NTrP(dudu1)N]\displaystyle=\frac{(-1)^{N}}{4^{N}}\big{[}\operatorname{Tr}(P\mathop{}\mathopen{}\mathrm{d}u^{-1}\mathop{}\mathopen{}\mathrm{d}u)^{N}-\operatorname{Tr}(P\mathop{}\mathopen{}\mathrm{d}u\mathop{}\mathopen{}\mathrm{d}u^{-1})^{N}\big{]}=\frac{(-1)^{N}}{4^{N}}\big{[}\operatorname{Tr}P(\mathop{}\mathopen{}\mathrm{d}u^{-1}\mathop{}\mathopen{}\mathrm{d}u)^{N}-\operatorname{Tr}P(\mathop{}\mathopen{}\mathrm{d}u\mathop{}\mathopen{}\mathrm{d}u^{-1})^{N}\big{]}
=(1)N4N[TrI+F2(du1du)NTrI+F2(dudu1)N]\displaystyle=\frac{(-1)^{N}}{4^{N}}\bigg{[}\operatorname{Tr}\tfrac{\mathrm{I}+F}{2}(\mathop{}\mathopen{}\mathrm{d}u^{-1}\mathop{}\mathopen{}\mathrm{d}u)^{N}-\operatorname{Tr}\tfrac{\mathrm{I}+F}{2}(\mathop{}\mathopen{}\mathrm{d}u\mathop{}\mathopen{}\mathrm{d}u^{-1})^{N}\bigg{]}
=(1)N24n[Tr(du1du)N+TrF(du1du)NTr(dudu1)NTrF(dudu1)N]\displaystyle=\frac{(-1)^{N}}{2\cdot 4^{n}}\bigg{[}\operatorname{Tr}(\mathop{}\mathopen{}\mathrm{d}u^{-1}\mathop{}\mathopen{}\mathrm{d}u)^{N}+\operatorname{Tr}F(\mathop{}\mathopen{}\mathrm{d}u^{-1}\mathop{}\mathopen{}\mathrm{d}u)^{N}-\operatorname{Tr}(\mathop{}\mathopen{}\mathrm{d}u\mathop{}\mathopen{}\mathrm{d}u^{-1})^{N}-\operatorname{Tr}F(\mathop{}\mathopen{}\mathrm{d}u\mathop{}\mathopen{}\mathrm{d}u^{-1})^{N}\bigg{]}
=(1)N4NTr(F(du1du)N)=(1)N4NTr(F[F,u1][F,u][F,u1]2N=n+1)\displaystyle=\frac{(-1)^{N}}{4^{N}}\operatorname{Tr}\big{(}F(\mathop{}\mathopen{}\mathrm{d}u^{-1}\mathop{}\mathopen{}\mathrm{d}u)^{N}\big{)}=\frac{(-1)^{N}}{4^{N}}\operatorname{Tr}\big{(}F\underbrace{[F,u^{-1}][F,u]\cdots[F,u^{-1}]}_{2N=n+1}\big{)}
=(6.1)(1)n+124n+12cnChnF(u1,u,,u1,u).\displaystyle\overset{\eqref{Chern-odd}}{=}\frac{(-1)^{\frac{n+1}{2}}}{4^{\frac{n+1}{2}}c_{n}}\mathrm{Ch}_{n}^{F}(u^{-1},u,\ldots,u^{-1},u).

   

Now, we describe the pairing (4.6) in the even case. The proof is left to the reader as an easy exercise.

Theorem 6.7 (even case)

Let (X,F,γ)(X,F,\gamma) be a qq-summable even Banach Fredholm module over the algebra 𝒜\mathcal{A} for some q>0q>0 with F2=IdF^{2}=\mathrm{Id}. Consider a class \mathscr{B} of Banach spaces stable under countable sums and containing the zero space. For any element [e][e] of the group K0(𝒜)\mathrm{K}_{0}(\mathcal{A}) and any even integer nqn\geqslant q, we have

(6.7) [e],[(X,F,γ)]K0(𝒜),K0(𝒜,)=(1)n22cnChnF(e,e,,e).\big{\langle}[e],[(X,F,\gamma)]\big{\rangle}_{\mathrm{K}_{0}(\mathcal{A}),\mathrm{K}^{0}(\mathcal{A},\mathscr{B})}=\frac{(-1)^{\frac{n}{2}}}{2c_{n}}\mathrm{Ch}_{n}^{F}(e,e,\ldots,e).

7 From Banach spectral triples to Banach Fredhlom modules

The following definition is extracted of [ArK22].

Definition 7.1 (Lipschitz algebra)

Consider a triple (𝒜,X,D)(\mathcal{A},X,D) constituted of the following data: a Banach space XX, a closed unbounded operator DD on XX with dense domain domDX\operatorname{dom}D\subset X, and an algebra 𝒜\mathcal{A} endowed with a homomorphism π:𝒜B(X)\pi\colon\mathcal{A}\to\mathrm{B}(X). In this case, we define the Lipschitz algebra

(7.1) LipD(𝒜)=def{a𝒜:π(a)domDdomD and the unbounded operator\displaystyle\operatorname{\mathrm{Lip}}_{D}(\mathcal{A})\overset{\mathrm{def}}{=}\big{\{}a\in\mathcal{A}:\pi(a)\cdot\operatorname{dom}D\subset\operatorname{dom}D\text{ and the unbounded operator }
[D,π(a)]:domDXX extends to an element of B(X)}.\displaystyle\qquad\qquad[D,\pi(a)]\colon\operatorname{dom}D\subset X\to X\text{ extends to an element of }\mathrm{B}(X)\big{\}}.

We refer to [ArK22, Section 5.7] for some properties of LipD(𝒜)\operatorname{\mathrm{Lip}}_{D}(\mathcal{A}). The following definition is essentially [ArK22, Definition 5.10 p. 218].

Definition 7.2 (compact Banach spectral triple)

Consider a triple (𝒜,X,D)(\mathcal{A},X,D) constituted of the following data: a reflexive1110. It may perhaps be possible to replace the reflexivity by an assumption of weak compactness, see [HvNVW18, p. 361]. Banach space XX, a closed unbounded bisectorial operator DD on XX with dense domain domDX\operatorname{dom}D\subset X, and a Banach algebra 𝒜\mathcal{A} equipped with a homomorphism π:𝒜B(X)\pi\colon\mathcal{A}\to\mathrm{B}(X). We say that (𝒜,X,D)(\mathcal{A},X,D) is a possibly kernel-degenerate compact Banach spectral triple if

  1. 1.

    DD admits a bounded H(Σωbi)\mathrm{H}^{\infty}(\Sigma^{\mathrm{bi}}_{\omega}) functional calculus on a bisector Σωbi\Sigma^{\mathrm{bi}}_{\omega}.

  2. 2.

    D1D^{-1} is a compact operator on RanD¯\overline{\operatorname{Ran}D}.

  3. 3.

    The subset LipD(A)\operatorname{\mathrm{Lip}}_{D}(A) is dense in 𝒜\mathcal{A}.

If KerD\operatorname{Ker}D is in addition finite-dimensional, we say that (𝒜,X,D)(\mathcal{A},X,D) is a compact Banach spectral triple.

In this situation, we have by [HvNVW18, p. 448] a direct sum decomposition X=RanD¯KerDX=\overline{\operatorname{Ran}D}\oplus\operatorname{Ker}D. Now, we introduce of summability similar to the hilbertian context. Recall that the space Sappq(X)S^{q}_{\mathrm{app}}(X) is defined in (5.4). The operator |D|1|D|^{-1} is well-defined on RanD¯\overline{\operatorname{Ran}D}. Furthermore, we can extend it by letting |D|1=0|D|^{-1}=0 on KerD\operatorname{Ker}D.

Definition 7.3 (summability)

Suppose that 1<q<1<q<\infty. We say that a possibly kernel-degenerate Banach spectral triple (𝒜,X,D)(\mathcal{A},X,D) is qq-summable if the operator |D|1|D|^{-1} is bounded and belongs to the space Sappq(X)S^{q}_{\mathrm{app}}(X).

The following definition is essentially [ArK22, Definition 5.11 p. 2123]. Compare to [CGRS14, Definition 2.1 p. 33]. We can also introduce variants for algebras which are not Banach algebras. These extensions are left to the reader.

Definition 7.4 (locally compact Banach spectral triple)

Consider a triple (𝒜,X,D)(\mathcal{A},X,D) constituted of the following data: a reflexive Banach space XX, a closed unbounded bisectorial operator DD on XX with dense domain domDX\operatorname{dom}D\subset X, and a Banach algebra 𝒜\mathcal{A} equipped with a homomorphism π:𝒜B(X)\pi\colon\mathcal{A}\to\mathrm{B}(X). We say that (𝒜,X,D)(\mathcal{A},X,D) is a possibly kernel-degenerate locally compact Banach spectral triple if

  1. 1.

    DD admits a bounded H(Σωbi)\mathrm{H}^{\infty}(\Sigma^{\mathrm{bi}}_{\omega}) functional calculus on a bisector Σωbi\Sigma^{\mathrm{bi}}_{\omega}.

  2. 2.

    π(a)|D|1|RanD¯\pi(a)|D|^{-1}|_{\overline{\operatorname{Ran}D}} is a compact operator on RanD¯X\overline{\operatorname{Ran}D}\to X.

  3. 3.

    The subset LipD(A)\operatorname{\mathrm{Lip}}_{D}(A) is dense in 𝒜\mathcal{A}.

Strong convex compactness property

Recall that compactness is preserved under strong integrals by [Voi92, Theorem 1.3 p. 260]. More precisely, for any measure space (Ω,μ)(\Omega,\mu) and any strongly measurable function T:Ω𝒦(X,Y)T\colon\Omega\to\mathcal{K}(X,Y), with values in the space 𝒦(X,Y)\mathcal{K}(X,Y) of compact operators from XX into YY, such that ΩT(ω)dμ(ω)\int_{\Omega}\left\|T(\omega)\right\|\mathop{}\mathopen{}\mathrm{d}\mu(\omega) is finite, the strong integral ΩT(ω)dμ(ω)\int_{\Omega}T(\omega)\mathop{}\mathopen{}\mathrm{d}\mu(\omega) is a compact operator.

Fractional powers

We need some background on fractional powers for a sectorial operator AA acting on a Banach space XX. If 0<Reα<10<\operatorname{Re}\alpha<1, we have by [Haa06, Proposition 3.2.1 p. 70] or [HvNVW23, p. 449]

(7.2) Aα=sinπαπ0tα(t+A)1xdt,xRanA.A^{-\alpha}=\frac{\sin\pi\alpha}{\pi}\int_{0}^{\infty}t^{-\alpha}(t+A)^{-1}x\mathop{}\mathopen{}\mathrm{d}t,\quad x\in\operatorname{Ran}A.
Proposition 7.5

Let (𝒜,X,D)(\mathcal{A},X,D) be a compact Banach spectral triple over an algebra 𝒜\mathcal{A}. Then (X,signD)(X,\operatorname{\mathrm{sign}}D) is a Banach Fredholm module over 𝒜\mathcal{A} (where XX is endowed with the same homomorphism π\pi).

Proof : We let F=defsignDF\overset{\mathrm{def}}{=}\operatorname{\mathrm{sign}}D. Since F2IF^{2}-I vanishes on RanD¯\overline{\operatorname{Ran}D} by construction, the first requirement in the Definition 3.1 of a Banach Fredholm module hold true.

The operator DD is bisectorial. Consequently, by [HvNVW18, Proposition 10.6.2 p. 448] the operator D2D^{2} is sectorial. Moreover, it admits a bounded H(Σθ)\mathrm{H}^{\infty}(\Sigma_{\theta}) functional calculus by [HvNVW18]. Recall that the functional calculus of DD and D2D^{2} are compatible. So we have

(7.3) |D|1x=(D2)12x=(7.2)1π0(t+D2)1xdtt,xRanD.|D|^{-1}x=(D^{2})^{-\frac{1}{2}}x\overset{\eqref{for-frac-powers}}{=}\frac{1}{\pi}\int_{0}^{\infty}(t+D^{2})^{-1}x\frac{\mathop{}\mathopen{}\mathrm{d}t}{\sqrt{t}},\quad x\in\operatorname{Ran}D.

The end of the proof is similar to the hilbertian one using that the compactness is preserved under strong integrals, Proposition 2.3 and [Haa06, Example 2.2.5 p. 29].    

Remark 7.6

A more simple proof can be given with the additional assumption that [|D|,a][|D|,a] defines a bounded operator on XX for all a𝒜a\in\mathcal{A}. Indeed, Let a𝒜a\in\mathcal{A}. Note that

(7.4) [signD,a]=D|D|1aaD|D|1=(DaaD)|D|1+D(|D|1aa|D|1)\displaystyle\left[\operatorname{\mathrm{sign}}D,a\right]=D|D|^{-1}a-aD|D|^{-1}=(Da-aD)|D|^{-1}+D(|D|^{-1}a-a|D|^{-1})
=[D,a]|D|1+D[|D|1,a]=[D,a]|D|1D|D|1[|D|,a]|D|1\displaystyle=[D,a]|D|^{-1}+D[|D|^{-1},a]=[D,a]|D|^{-1}-D|D|^{-1}[|D|,a]|D|^{-1}
=([D,a]sign(D)[|D|,a]|=defR)|D|1.\displaystyle=\big{(}\underbrace{[D,a]-\operatorname{\mathrm{sign}}(D)[|D|,a]|}_{\overset{\mathrm{def}}{=}R}\big{)}|D|^{-1}.

This operator is compact. Let aπ(𝒜)¯a\in\overline{\pi(\mathcal{A})}. There exists a sequence (ak)(a_{k}) of elements of 𝒜\mathcal{A} such that π(ak)a\pi(a_{k})\to a. Each operator [signD,ak][\operatorname{\mathrm{sign}}D,a_{k}] is compact by the first part of the proof. We have [signD,ak][signD,a][\operatorname{\mathrm{sign}}D,a_{k}]\to\left[\operatorname{\mathrm{sign}}D,a\right] since the operator product is continuous in B(X)\mathrm{B}(X). We conclude that the operator [signD,a]\left[\operatorname{\mathrm{sign}}D,a\right] is compact.

Proposition 7.7

Suppose that (𝒜,X,D)(\mathcal{A},X,D) is a qq-summable Banach spectral triple. Assume that [|D|,a][|D|,a] is a well-defined bounded operator for any a𝒜a\in\mathcal{A}. Then (X,signD)(X,\operatorname{\mathrm{sign}}D) is a qq-summable Banach Fredholm module for 𝒜\mathcal{A}.

Proof : Let a𝒜a\in\mathcal{A}. Recall that

[signD,a]=(7.4)R|D|1.\left[\operatorname{\mathrm{sign}}D,a\right]\overset{\eqref{formula-simple}}{=}R|D|^{-1}.

By the ideal property (5.3), the operator R|D|1R|D|^{-1} belongs to the space Sappq(X)S^{q}_{\mathrm{app}}(X).    

8 Examples of triples and Banach Fredholm modules

8.1 The Dirac operator 1iddθ\frac{1}{\mathrm{i}}\frac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}\theta} on the space Lp(𝕋)\mathrm{L}^{p}(\mathbb{T}) and the periodic Hilbert transform

Suppose that 1<p<1<p<\infty. Here 𝕋={z:|z|=1}\mathbb{T}=\{z\in\mathbb{C}:|z|=1\} is the one-dimensional torus. The functions defined on 𝕋\mathbb{T} can be identified with periodic functions on \mathbb{R} with period 2π2\pi. In particular, we have

C(𝕋)={gC():f(θ+2π)=f(θ)}\mathrm{C}(\mathbb{T})=\{g\in\mathrm{C}(\mathbb{R}):f(\theta+2\pi)=f(\theta)\}

Recall that any multiplication operator Mf:Lp(𝕋)Lp(𝕋)M_{f}\colon\mathrm{L}^{p}(\mathbb{T})\to\mathrm{L}^{p}(\mathbb{T}), fgfgfg\mapsto fg by a function ff of L(𝕋)\mathrm{L}^{\infty}(\mathbb{T}) is bounded, see [EnN00, Proposition 4.10 p. 31]. Consequently, we can consider the homomorphism π:L(𝕋)B(Lp(𝕋))\pi\colon\mathrm{L}^{\infty}(\mathbb{T})\to\mathrm{B}(\mathrm{L}^{p}(\mathbb{T})), fMff\mapsto M_{f}. It is well-known [Kha13, Exercise 4.2.6 p. 194] [GVF01, p. 390] that (C(𝕋),L2(𝕋),1iddθ)(\mathrm{C}^{\infty}(\mathbb{T}),\mathrm{L}^{2}(\mathbb{T}),\frac{1}{\mathrm{i}}\frac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}\theta}) is a compact spectral triple, where the used homomorphism is the restriction of π\pi on the algebra C(𝕋)\mathrm{C}^{\infty}(\mathbb{T}). First, we prove an Lp\mathrm{L}^{p}-generalization of this classical fact. Here, we consider the closure of the unbounded operator ddθ:C(𝕋)Lp(𝕋)Lp(𝕋)\frac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}\theta}\colon\mathrm{C}^{\infty}(\mathbb{T})\subset\mathrm{L}^{p}(\mathbb{T})\to\mathrm{L}^{p}(\mathbb{T}). We denote again ddθ\frac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}\theta} its closure whose the domain domddθ\operatorname{dom}\frac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}\theta} is the Sobolev space

W1,p(𝕋)=def{fLp():f(θ+2π)=f(θ),f is absolutely continuous,fLp()}.\mathrm{W}^{1,p}(\mathbb{T})\overset{\mathrm{def}}{=}\{f\in\mathrm{L}^{p}(\mathbb{R}):f(\theta+2\pi)=f(\theta),f\text{ is absolutely continuous},f^{\prime}\in\mathrm{L}^{p}(\mathbb{R})\}.

Suppose that 1<p<1<p<\infty and α0\alpha\geqslant 0. We define the subspace

(8.1) Lαp(𝕋)=defdomΔpα2\mathrm{L}^{p}_{\alpha}(\mathbb{T})\overset{\mathrm{def}}{=}\operatorname{dom}\Delta^{\frac{\alpha}{2}}_{p}

of Lp(𝕋)\mathrm{L}^{p}(\mathbb{T}). If fLαp(𝕋)f\in\mathrm{L}^{p}_{\alpha}(\mathbb{T}), we will use the notation

(8.2) fLαp(𝕋)=def(Id+Δp)α2(f)Lp(𝕋).\left\|f\right\|_{\mathrm{L}^{p}_{\alpha}(\mathbb{T})}\overset{\mathrm{def}}{=}\big{\|}(\mathrm{Id}+\Delta_{p})^{\frac{\alpha}{2}}(f)\big{\|}_{\mathrm{L}^{p}(\mathbb{T})}.

Suppose 1<q<p<1<q<p<\infty and α>0\alpha>0 such that αd1q1p\frac{\alpha}{d}\geqslant\frac{1}{q}-\frac{1}{p}. We will use the Sobolev inequality

(8.3) fLp(𝕋d)fLαq(𝕋d).\left\|f\right\|_{\mathrm{L}^{p}(\mathbb{T}^{d})}\lesssim\left\|f\right\|_{\mathrm{L}^{q}_{\alpha}(\mathbb{T}^{d})}.

of [BEO13, Corollary 1.2], which gives a map J:Lαq(𝕋d)Lp(𝕋d)J\colon\mathrm{L}^{q}_{\alpha}(\mathbb{T}^{d})\to\mathrm{L}^{p}(\mathbb{T}^{d}). Moreover, it is known [Kon86, Theorem p. 187] that if αd>1q1p\frac{\alpha}{d}>\frac{1}{q}-\frac{1}{p}

(8.4) an(J)1nαd,n0.a_{n}(J)\lesssim\frac{1}{n^{\frac{\alpha}{d}}},\quad n\geqslant 0.

The main point of the following result is the new connection between Sobolev embedding and summability in the proof.

Theorem 8.1

Suppose that 1<p<1<p<\infty. The triple (C(𝕋),Lp(𝕋),1iddθ)(\mathrm{C}^{\infty}(\mathbb{T}),\mathrm{L}^{p}(\mathbb{T}),\frac{1}{\mathrm{i}}\frac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}\theta}) is a Banach compact spectral triple which is qq-summable for any q>1q>1.

Proof : Note that it is well-known that the operator ddθ\frac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}\theta} generate a strongly continuous group of operators acting on Lp(𝕋)\mathrm{L}^{p}(\mathbb{T}), namely the group of translations. By Example 2.4 and Example 2.5, we deduce that the unbounded operator 1iddθ\frac{1}{\mathrm{i}}\frac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}\theta} is bisectorial of angle 0 and admits a bounded H(Σωbi)\mathrm{H}^{\infty}(\Sigma_{\omega}^{\mathrm{bi}}) functional calculus for any angle ω>0\omega>0 on the Banach space Lp(𝕋)\mathrm{L}^{p}(\mathbb{T}). For any function fC(𝕋)f\in\mathrm{C}^{\infty}(\mathbb{T}), an elementary computation reveals that MfdomddθdomddθM_{f}\cdot\operatorname{dom}\tfrac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}\theta}\subset\operatorname{dom}\tfrac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}\theta} and that

(8.5) [1iddθ,f]=1iMf.\big{[}\tfrac{1}{\mathrm{i}}\tfrac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}\theta},f\big{]}=\frac{1}{\mathrm{i}}M_{f^{\prime}}.

So the commutator [1iddθ,f][\frac{1}{\mathrm{i}}\tfrac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}\theta},f] defines a bounded operator on the Banach space Lp(𝕋)\mathrm{L}^{p}(\mathbb{T}). Finally, it is stated (without proof; the end of the proof below is stronger than this fact) in [Kat76, Example 6.31 p. 187] the operator 1iddθ\frac{1}{\mathrm{i}}\tfrac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}\theta} has compact resolvent. So Definition 7.4 is satisfied.

Suppose that 1<q<p<21<q<p<2. Let α>0\alpha>0. For any function fLp(𝕋)f\in\mathrm{L}^{p}(\mathbb{T}), we have

(Id+Δ)α2(f)Lαq(𝕋)=(8.2)(Id+Δ)α2(Id+Δ)α2fLq(𝕋)=fLq(𝕋)xLp(𝕋).\left\|(\mathrm{Id}+\Delta)^{-\frac{\alpha}{2}}(f)\right\|_{\mathrm{L}^{q}_{\alpha}(\mathbb{T})}\overset{\eqref{Def-Lpalpha-bis}}{=}\left\|(\mathrm{Id}+\Delta)^{\frac{\alpha}{2}}(\mathrm{Id}+\Delta)^{-\frac{\alpha}{2}}f\right\|_{\mathrm{L}^{q}(\mathbb{T})}=\left\|f\right\|_{\mathrm{L}^{q}(\mathbb{T})}\lesssim\left\|x\right\|_{\mathrm{L}^{p}(\mathbb{T})}.

If α>1q1p\alpha>\frac{1}{q}-\frac{1}{p}, using the Sobolev inequality (8.3), we deduce the factorization

(8.6) (Id+Δ)α2:Lp(𝕋)(Id+Δ)α2Lαq(𝕋)𝐽Lp(𝕋).(\mathrm{Id}+\Delta)^{-\frac{\alpha}{2}}\colon\mathrm{L}^{p}(\mathbb{T})\xrightarrow{(\mathrm{Id}+\Delta)^{-\frac{\alpha}{2}}}\mathrm{L}^{q}_{\alpha}(\mathbb{T})\xrightarrow{J}\mathrm{L}^{p}(\mathbb{T}).

With [Kon86, Theorem p. 187], we obtain for any integer n1n\geqslant 1

an((Id+Δ)α2)=(8.6)an(J(Id+Δ)α2)(5.1)(Id+Δ)α2an(J)(8.4)1nα.\displaystyle a_{n}((\mathrm{Id}+\Delta)^{-\frac{\alpha}{2}})\overset{\eqref{facto-1}}{=}a_{n}(J(\mathrm{Id}+\Delta)^{-\frac{\alpha}{2}})\overset{\eqref{majo-sn-2}}{\leqslant}\left\|(\mathrm{Id}+\Delta)^{-\frac{\alpha}{2}}\right\|a_{n}(J)\overset{\eqref{sn-de-Sobolev}}{\lesssim}\frac{1}{n^{\alpha}}.

The series an((Id+Δ)α2)\sum a_{n}((\mathrm{Id}+\Delta)^{-\frac{\alpha}{2}}) converges if α>1\alpha>1. We deduce that the operator (Id+Δ)α2:Lp(𝕋)Lp(𝕋)(\mathrm{Id}+\Delta)^{-\frac{\alpha}{2}}\colon\mathrm{L}^{p}(\mathbb{T})\to\mathrm{L}^{p}(\mathbb{T}) belongs to the space Sapp1(Lp(𝕋))S^{1}_{\mathrm{app}}(\mathrm{L}^{p}(\mathbb{T})). Thus the triple (C(𝕋),Lp(),1iddx)(\mathrm{C}^{\infty}(\mathbb{T}),\mathrm{L}^{p}(\mathbb{R}),\frac{1}{\mathrm{i}}\frac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}x}) is qq-summable for any q>1q>1. If p>2p>2, we reason by duality.    

Remark 8.2

Suppose that 1<p<1<p<\infty. In this remark, we sketch a proof explaing why the previous result is optimal. If p=2p=2 it is well-known. The Banach compact spectral triple (C(𝕋),Lp(𝕋),1iddθ)(\mathrm{C}^{\infty}(\mathbb{T}),\mathrm{L}^{p}(\mathbb{T}),\frac{1}{\mathrm{i}}\frac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}\theta}) is not 11-summable. Indeed, by [AmG16, Remark B.7], the spectrum of the Dirac operator D=def1iddθD\overset{\mathrm{def}}{=}\frac{1}{\mathrm{i}}\frac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}\theta} is independent of pp and the Laplacian has the same property by [Dav89, Theorem 1.6.3 p. 36]. In the case p=2p=2, it is known that SpD=\operatorname{Sp}D=\mathbb{Z}. Using Weyl’s inequality (5.7) and a ¡¡bisectorial version¿¿ (left to the reader) of the spectral theorem of [Haa06, p. 56] [Haa05], we see that for any μSpD\mu\not\in\operatorname{Sp}D

n1|μn|=(λn(Rμ(D)))1(5.7)Rμ(D)Sweyl1(Lp(𝕋))Rμ(D)Sapp1(Lp(𝕋)),\displaystyle\sum_{n\in\mathbb{Z}}\frac{1}{|\mu-n|}=\left\|\big{(}\lambda_{n}(R_{\mu}(D))\big{)}\right\|_{\ell^{1}}\overset{\eqref{Weyl-inequality}}{\lesssim}\left\|R_{\mu}(D)\right\|_{S^{1}_{\mathrm{weyl}}(\mathrm{L}^{p}(\mathbb{T}))}\leqslant\left\|R_{\mu}(D)\right\|_{S^{1}_{\mathrm{app}}(\mathrm{L}^{p}(\mathbb{T}))},

where we use the resolvent operator Rμ(D)R_{\mu}(D). We conclude that the triple is not 11-summable according to Definition 7.3.

Recall that the periodic Hilbert transform :Lp(𝕋)Lp(𝕋)\mathcal{H}\colon\mathrm{L}^{p}(\mathbb{T})\to\mathrm{L}^{p}(\mathbb{T}) is defined by the principal value [Kin09a, (3.285) p. 132] [Cas22, p. 181]

(8.7) (f)(θ)=def12πlimε0+ε|s|πf(θs)cot(s2)ds.(\mathcal{H}f)(\theta)\overset{\mathrm{def}}{=}\frac{1}{2\pi}\lim_{\varepsilon\to 0^{+}}\int_{\varepsilon\leqslant|s|\leqslant\pi}f(\theta-s)\cot(\tfrac{s}{2})\mathop{}\mathopen{}\mathrm{d}s.

and can be seen as a Fourier multiplier with symbol isign(n)-\mathrm{i}\,\operatorname{\mathrm{sign}}(n), i.e. we have

(f)^(n)=isign(n)f^(n),n,\widehat{\mathcal{H}(f)}(n)=-\mathrm{i}\operatorname{\mathrm{sign}}(n)\hat{f}(n),\quad n\in\mathbb{Z},

where sign(0)=def0\operatorname{\mathrm{sign}}(0)\overset{\mathrm{def}}{=}0. The boundedness of the periodic Hilbert transform \mathcal{H} on the space Lp(𝕋)\mathrm{L}^{p}(\mathbb{T}) is proved in [Kin09a, Section 6.17] or [HvNVW16, Proposition 5.2.5 p. 391] (by transference from the Hilbert transform). Another proof is provided by the next result. Now, we determine the index pairing of the associated Banach Fredholm module by Proposition 7.5.

Recall that the winding number of a a function f:𝕋{0}f\colon\mathbb{T}\to\mathbb{C}-\{0\} is the number of turns of the point f(eit)f(e^{\mathrm{i}t}) around the origin when tt runs from 0 to 2π2\pi. More precisely, consider a continuous branch of the argument argf\arg_{f} of the function [0,2π]{0}[0,2\pi]\to\mathbb{C}-\{0\}, tf(eit)t\mapsto f(\mathrm{e}^{\mathrm{i}t}), i.e. argf\arg_{f} is continuous on [0,2π][0,2\pi] and

f(eit)|f(eit)|=eiargf(t),t[0,2π]\frac{f(\mathrm{e}^{\mathrm{i}t})}{|f(\mathrm{e}^{\mathrm{i}t})|}=\mathrm{e}^{\mathrm{i}\arg_{f}(t)},\quad t\in[0,2\pi]

The winding number of ff is defined by

(8.8) windf=def12π(argf(2π)argf(0)).\operatorname{wind}f\overset{\mathrm{def}}{=}\frac{1}{2\pi}(\arg_{f}(2\pi)-\arg_{f}(0)).

It is worth noting that if fC1(𝕋)f\in\mathrm{C}^{1}(\mathbb{T}) and does not vanish on 𝕋\mathbb{T} then

windf=12πi02πf(eiθ)f(eiθ)dθ.\operatorname{wind}f=\frac{1}{2\pi\mathrm{i}}\int_{0}^{2\pi}\frac{f^{\prime}(\mathrm{e}^{\mathrm{i}\theta})}{f(\mathrm{e}^{\mathrm{i}\theta})}\mathop{}\mathopen{}\mathrm{d}\theta.

Now, we describe the index pairing. The case p=2p=2 is folklore.

Proposition 8.3

Suppose that 1<p<1<p<\infty. The kernel-degenerate odd Banach Fredholm module over the algebra C(𝕋)\mathrm{C}(\mathbb{T}) associated to the Banach compact spectral triple (C(𝕋),Lp(𝕋),1iddθ)(\mathrm{C}^{\infty}(\mathbb{T}),\mathrm{L}^{p}(\mathbb{T}),\frac{1}{\mathrm{i}}\frac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}\theta}) is (Lp(𝕋),i)(\mathrm{L}^{p}(\mathbb{T}),\mathrm{i}\mathcal{H}). For any function fC(𝕋)f\in\mathrm{C}(\mathbb{T}) which does not vanish on 𝕋\mathbb{T}, we have

(8.9) [f],(Lp(𝕋),i)K1(C(𝕋)),K1(C(𝕋),p)=windf.\big{\langle}[f],(\mathrm{L}^{p}(\mathbb{T}),\mathrm{i}\mathcal{H})\big{\rangle}_{\mathrm{K}_{1}(\mathrm{C}(\mathbb{T})),\mathrm{K}^{1}(\mathrm{C}(\mathbb{T}),\mathscr{L}^{p})}=-\operatorname{wind}f.

Proof : The first sentence is obvious. We let F=defiF\overset{\mathrm{def}}{=}\mathrm{i}\mathcal{H}. Note that the map Q:Lp(𝕋)Lp(𝕋)Q\colon\mathrm{L}^{p}(\mathbb{T})\to\mathrm{L}^{p}(\mathbb{T}), einθδn=0\mathrm{e}^{\mathrm{i}n\theta}\to\delta_{n=0} (which maps a function on the constant term in its Fourier series) is a finite-rank bounded map. So F=defF+QF^{\prime}\overset{\mathrm{def}}{=}F+Q is clearly a Banach Fredholm module since (F)2=IdLp(𝕋)(F^{\prime})^{2}=\mathrm{Id}_{\mathrm{L}^{p}(\mathbb{T})} and a compact perturbation of FF. Moreover, observe that the map P=defId+F2P\overset{\mathrm{def}}{=}\frac{\mathrm{Id}+F^{\prime}}{2} identifies to the Riesz projection Lp(𝕋)Lp(𝕋)\mathrm{L}^{p}(\mathbb{T})\to\mathrm{L}^{p}(\mathbb{T}), einθδn0einθ\mathrm{e}^{\mathrm{i}n\theta}\mapsto\delta_{n\geqslant 0}\mathrm{e}^{\mathrm{i}n\theta}. The range of this map is the Hardy space Hp(𝕋)\mathrm{H}^{p}(\mathbb{T}). Let fC(𝕋)f\in\mathrm{C}(\mathbb{T}) which does not vanish on 𝕋\mathbb{T}. This means that ff is an invertible element of the unital algebra C(𝕋)\mathrm{C}(\mathbb{T}). The operator PMfP:P(Lp(𝕋))P(Lp(𝕋))PM_{f}P\colon P(\mathrm{L}^{p}(\mathbb{T}))\to P(\mathrm{L}^{p}(\mathbb{T})) of (4.1) (with n=1n=1) identifies to the Toeplitz operator Tf:Hp(𝕋)Hp(𝕋)T_{f}\colon\mathrm{H}^{p}(\mathbb{T})\to\mathrm{H}^{p}(\mathbb{T}), gP(fg)g\mapsto P(fg) with symbol fL(𝕋)f\in\mathrm{L}^{\infty}(\mathbb{T}), where we identify the map PP with its corestriction P|Hp(𝕋)P|^{\mathrm{H}^{p}(\mathbb{T})}. So the operator TfT_{f} is Fredholm. Using the classical Gohberg-Krein index theorem [BoS06, Theorem 2.42 p. 74], we deduce that

[f],(Lp(𝕋),i)K1(C(𝕋)),K1(C(𝕋),p)=(4.4)IndexPMfP=IndexTf=windf.\displaystyle\big{\langle}[f],(\mathrm{L}^{p}(\mathbb{T}),\mathrm{i}\mathcal{H})\big{\rangle}_{\mathrm{K}_{1}(\mathrm{C}(\mathbb{T})),\mathrm{K}^{1}(\mathrm{C}(\mathbb{T}),\mathscr{L}^{p})}\overset{\eqref{pairing-odd-2}}{=}\operatorname{Index}PM_{f}P=\operatorname{Index}T_{f}=-\operatorname{wind}f.

   

Recall that the first group of K\mathrm{K}-theory is given by K1(C(𝕋))=\mathrm{K}_{1}(\mathrm{C}(\mathbb{T}))=\mathbb{Z}. By the way, we refer to [ScS23] for a nice study of the K-groups of spheres.

Remark 8.4

It is folklore that for any function fL(𝕋)f\in\mathrm{L}^{\infty}(\mathbb{T}) the commutator [F,f][F,f] is compact on Lp(𝕋)\mathrm{L}^{p}(\mathbb{T}) if and only if ff belongs to the space VMO(𝕋)\mathrm{VMO}(\mathbb{T}) of functions of vanishing mean oscillation. In the case p=2p=2, it is also known (essentially in [JaW82]) that the commutator [F,f][F,f] belongs to the Schatten space Sq(L2(𝕋))S^{q}(\mathrm{L}^{2}(\mathbb{T})) if and only if ff belongs to the Besov space Bq,q1q(𝕋)\mathrm{B}^{\frac{1}{q}}_{q,q}(\mathbb{T}).

Remark 8.5

It is easy to compute the Chern character in the case 1<p<1<p<\infty. The case p=2p=2 is folklore.

8.2 The Dirac operator 1iddx\frac{1}{\mathrm{i}}\frac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}x} on the space Lp()\mathrm{L}^{p}(\mathbb{R}) and the Hilbert transform

Suppose that 1<p<1<p<\infty. Recall that any multiplication operator Mf:Lp()Lp()M_{f}\colon\mathrm{L}^{p}(\mathbb{R})\to\mathrm{L}^{p}(\mathbb{R}), fgfgfg\mapsto fg by a function ff of L()\mathrm{L}^{\infty}(\mathbb{R}) is bounded, see [EnN00, Proposition 4.10 p. 31]. Consequently, we can consider the homomorphism π:L()B(Lp())\pi\colon\mathrm{L}^{\infty}(\mathbb{R})\to\mathrm{B}(\mathrm{L}^{p}(\mathbb{R})), fMff\mapsto M_{f}. It is well-known [Ren04, Corollary 14 p. 92] that (C0(),L2(),1iddx)(\mathrm{C}^{\infty}_{0}(\mathbb{R}),\mathrm{L}^{2}(\mathbb{R}),\frac{1}{\mathrm{i}}\frac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}x}) is a locally compact spectral triple, where the used homomorphism is the restriction of π\pi on the algebra C0()\mathrm{C}^{\infty}_{0}(\mathbb{R}). First, we prove a weaker Lp\mathrm{L}^{p}-variant of this classical fact. Here, we consider the closure of the unbounded operator ddx:C0()Lp()Lp()\frac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}x}\colon\mathrm{C}^{\infty}_{0}(\mathbb{R})\subset\mathrm{L}^{p}(\mathbb{R})\to\mathrm{L}^{p}(\mathbb{R}). We denote again ddx\frac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}x} its closure whose the domain domddx\operatorname{dom}\frac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}x} is the Sobolev space

W1,p()=def{fLp():f is absolutely continuous,fLp()}.\mathrm{W}^{1,p}(\mathbb{R})\overset{\mathrm{def}}{=}\{f\in\mathrm{L}^{p}(\mathbb{R}):f\text{ is absolutely continuous},f^{\prime}\in\mathrm{L}^{p}(\mathbb{R})\}.
Proposition 8.6

Suppose that 2p<2\leqslant p<\infty. Consider the triple (C0(),Lp(),1iddx)(\mathrm{C}_{0}^{\infty}(\mathbb{R}),\mathrm{L}^{p}(\mathbb{R}),\frac{1}{\mathrm{i}}\frac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}x}). The operator 1iddθ\frac{1}{\mathrm{i}}\frac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}\theta} is bisectorial of angle 0 and admits a bounded H(Σωbi)\mathrm{H}^{\infty}(\Sigma_{\omega}^{\mathrm{bi}}) functional calculus for any angle ω>0\omega>0 and the commutator [1iddx,f]\big{[}\tfrac{1}{\mathrm{i}}\tfrac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}x},f\big{]} is a well-defined bounded operator for any function fC0()f\in\mathrm{C}^{\infty}_{0}(\mathbb{R}).

Proof : Note that it is well-known [EnN00, Proposition 1 p. 66] that the operator ddx\frac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}x} generate a strongly continuous group of operators acting on Lp()\mathrm{L}^{p}(\mathbb{R}), namely the group of translations. By Example 2.4 and Example 2.5, we deduce that the unbounded operator 1iddθ\frac{1}{\mathrm{i}}\frac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}\theta} is bisectorial of angle 0 and admits a bounded H(Σωbi)\mathrm{H}^{\infty}(\Sigma_{\omega}^{\mathrm{bi}}) functional calculus for any angle ω>0\omega>0 on the Banach space Lp()\mathrm{L}^{p}(\mathbb{R}). For any function fCc()f\in\mathrm{C}_{c}^{\infty}(\mathbb{R}), an elementary computation reveals that MfdomddxdomddxM_{f}\cdot\operatorname{dom}\tfrac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}x}\subset\operatorname{dom}\tfrac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}x} and that

(8.10) [1iddx,f]=1iMf.\big{[}\tfrac{1}{\mathrm{i}}\tfrac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}x},f\big{]}=\frac{1}{\mathrm{i}}M_{f^{\prime}}.

So [1iddx,f][\frac{1}{\mathrm{i}}\tfrac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}x},f] defines a bounded operator on the Banach space Lp()\mathrm{L}^{p}(\mathbb{R}).    

Now, we consider the Hilbert transform :Lp()Lp()\mathcal{H}\colon\mathrm{L}^{p}(\mathbb{R})\to\mathrm{L}^{p}(\mathbb{R}). This transformation is a bounded operator by [HvNVW16, Theorem 5.1.1 p. 374] defined by the principal value

(8.11) (f)(x)=def1πp.v.f(y)xydy,f𝒮(), for almost every x(\mathcal{H}f)(x)\overset{\mathrm{def}}{=}\frac{1}{\pi}\operatorname{p.v.}\int_{\mathbb{R}}\frac{f(y)}{x-y}\mathop{}\mathopen{}\mathrm{d}y,\quad f\in\mathcal{S}(\mathbb{R}),\text{ for almost every }x\in\mathbb{R}

and can be seen by [HvNVW16, Proposition 5.2.2 p. 389] as a Fourier multiplier with symbol isign(ξ)-\mathrm{i}\,\operatorname{\mathrm{sign}}(\xi), i.e. we have

(8.12) (f)^(ξ)=isign(ξ)f^(ξ),f𝒮(),ξ.\widehat{\mathcal{H}(f)}(\xi)=-\mathrm{i}\,\operatorname{\mathrm{sign}}(\xi)\hat{f}(\xi),\quad f\in\mathcal{S}(\mathbb{R}),\xi\in\mathbb{R}.

Note that it was observed in [Con94, p. 314] that i\mathrm{i}\mathcal{H} is a Fredholm module on the Hilbert space L2()\mathrm{L}^{2}(\mathbb{R}). Note that the commutator is defined by (1.2)

For the computation of the index pairing, we need some information on Hardy spaces. Recall the notation +=def{z:Imz>0}\mathbb{C}_{+}\overset{\mathrm{def}}{=}\{z\in\mathbb{C}:\mathrm{Im}z>0\}. Following [Nik02, Definition 6.3.2 p. 145], we denote by Hp(+)\mathrm{H}^{p}(\mathbb{C}_{+}) the space of all functions FF which are analytic in the upper half plane +\mathbb{C}_{+} such that

FHp(+)=defsupy>0(|F(x+iy)|pdx)1p<,\left\|F\right\|_{\mathrm{H}^{p}(\mathbb{C}_{+})}\overset{\mathrm{def}}{=}\sup_{y>0}\left(\int_{\mathbb{R}}|F(x+\mathrm{i}y)|^{p}\mathop{}\mathopen{}\mathrm{d}x\right)^{\frac{1}{p}}<\infty,

and supz+|F(z)|<\sup_{z\in\mathbb{C}_{+}}|F(z)|<\infty for p=p=\infty. We also introduce the closed subspace

Hp()={fLp():f(t)tz¯dt=0 for all z+}\mathrm{H}^{p}(\mathbb{R})=\left\{f\in\mathrm{L}^{p}(\mathbb{R}):\int_{\mathbb{R}}\frac{f(t)}{t-\overline{z}}\mathop{}\mathopen{}\mathrm{d}t=0\text{ for all }z\in\mathbb{C}_{+}\right\}

of the Banach space Lp()\mathrm{L}^{p}(\mathbb{R}). Cauchy’s representation theorem in Hardy spaces, e.g. [Mas09, Chapter 13], can be formulated as follows. Consider some analytic function F:+F\colon\mathbb{C}_{+}\to\mathbb{C}. Then the following assertions are equivalent.

  • (i)

    The function FF belongs to the space Hp(+)\mathrm{H}^{p}(\mathbb{C}_{+}).

  • (ii)

    There exists a unique function fHp()f\in\mathrm{H}^{p}(\mathbb{R}) such that

    (8.13) F(z)=12πif(t)tzdt,z+.F(z)=\frac{1}{2\pi\mathrm{i}}\int_{\mathbb{R}}\frac{f(t)}{t-z}\mathop{}\mathopen{}\mathrm{d}t,\quad z\in\mathbb{C_{+}}.

In this case, we have

FHp(+)=fLp().\left\|F\right\|_{\mathrm{H}^{p}(\mathbb{C}_{+})}=\left\|f\right\|_{\mathrm{L}^{p}(\mathbb{R})}.

and the nontangential boundary function of FF is equal to ff, i.e. f(x)=limzx,z+F(z)f(x)=\lim_{z\to x,z\in\mathbb{C}_{+}}F(z). Consequently, we have an isometric isomorphism of the Hardy space Hp(+)\mathrm{H}^{p}(\mathbb{C}_{+}) onto the space Hp()\mathrm{H}^{p}(\mathbb{R}). Thus the Hardy space Hp(+)\mathrm{H}^{p}(\mathbb{C}_{+}) can be identified with a closed subspace of the Banach space Lp()\mathrm{L}^{p}(\mathbb{R}).

The Toeplitz operator Tg:Hp()Hp()T_{g}\colon\mathrm{H}^{p}(\mathbb{R})\to\mathrm{H}^{p}(\mathbb{R}) with symbol gL()g\in\mathrm{L}^{\infty}(\mathbb{R}) is defined by

(8.14) Tg(f)=defP+(gf),fHp().T_{g}(f)\overset{\mathrm{def}}{=}P_{+}(gf),\quad f\in\mathrm{H}^{p}(\mathbb{R}).

If f:{}f\colon\mathbb{R}\cup\{\infty\}\to\mathbb{C} is a continuous function then the image of ff in the complex plane is a closed curve Γ\Gamma. If

(8.15) infx{}|f(x)|>0,\inf\limits_{x\in\mathbb{R}\cup\{\infty\}}|f(x)|>0,

we define the winding number windf\operatorname{wind}f as the winding number of Γ\Gamma around the origin. The condition (8.15) is equivalent to say that ff is invertible in the algebra C({})\mathrm{C}(\mathbb{R}\cup\{\infty\}).

We will use the following result [Cam17, Theorem 6.2].

Theorem 8.7

Let fC({})f\in\mathrm{C}(\mathbb{R}\cup\{\infty\}). The operator Tf:Hp()Hp()T_{f}\colon\mathrm{H}^{p}(\mathbb{R})\to\mathrm{H}^{p}(\mathbb{R}) has closed range if and only if (8.15) is satisfied. In this case, the operator TfT_{f} is Fredholm and its Fredholm index is IndexTf=windf\operatorname{Index}T_{f}=-\operatorname{wind}f.

Remark 8.8

It is known that the condition (8.15) is equivalent to the existence of a Wiener-Hopf pp-factorization of ff. In this case, f=frkf+f=f_{-}r^{k}f_{+} with k=windfk=\operatorname{wind}f.

The following observation is new even if p=2p=2. Recall that K1(C0())=\mathrm{K}_{1}(\mathrm{C}_{0}(\mathbb{R}))=\mathbb{Z}, see e.g. [WeO93].

Proposition 8.9

Suppose that 1<p<1<p<\infty. The Banach odd Fredholm module over the algebra C0()\mathrm{C}_{0}(\mathbb{R}) associated to the triple (C0(),Lp(),1iddx)(\mathrm{C}^{\infty}_{0}(\mathbb{R}),\mathrm{L}^{p}(\mathbb{R}),\frac{1}{\mathrm{i}}\frac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}x}) is (Lp(),i)(\mathrm{L}^{p}(\mathbb{R}),\mathrm{i}\mathcal{H}). For any function fC0()f\in\mathrm{C}_{0}^{\infty}(\mathbb{R}) satisfying (8.15), we have

(8.16) [f],(Lp(),i)K1(C0()),K1(C0(),p)=windf,\big{\langle}[f],(\mathrm{L}^{p}(\mathbb{R}),\mathrm{i}\mathcal{H})\big{\rangle}_{\mathrm{K}_{1}(\mathrm{C}_{0}(\mathbb{R})),\mathrm{K}^{1}(\mathrm{C}_{0}(\mathbb{R}),\mathscr{L}^{p})}=-\operatorname{wind}f,

where is the winding number of ff.

Proof : We let F=defiF\overset{\mathrm{def}}{=}\mathrm{i}\mathcal{H}. We have F2=IdF^{2}=\mathrm{Id}. It is known [Uch78, Theorem 2 p. 17] that a function fL()f\in\mathrm{L}^{\infty}(\mathbb{R}) belongs to the space VMO()\mathrm{VMO}(\mathbb{R}) of functions of vanishing mean oscillation if and only if the commutator [F,Mf][F,M_{f}] is a compact operator acting on the Banach space Lp()\mathrm{L}^{p}(\mathbb{R}). We infer that (Lp(),F)(\mathrm{L}^{p}(\mathbb{R}),F) is a odd Fredhlom module over the algebra C0()\mathrm{C}_{0}(\mathbb{R}) of continuous functions that vanish at infinity, since C0()\mathrm{C}_{0}(\mathbb{R}) is a subspace of VMO()\mathrm{VMO}(\mathbb{R}).

Moreover, observe that the map P=defId+F2P\overset{\mathrm{def}}{=}\frac{\mathrm{Id}+F}{2} identifies to the bounded projection Lp()Lp()\mathrm{L}^{p}(\mathbb{R})\to\mathrm{L}^{p}(\mathbb{R}) on the subspace Hp()\mathrm{H}^{p}(\mathbb{R}). Let fC0()f\in\mathrm{C}_{0}(\mathbb{R}) satisfying (8.15). This means that ff is an invertible element of the unitization C(X{})\mathrm{C}(X\cup\{\infty\}) of the algebra C0(X)\mathrm{C}_{0}(X). The operator PMfP:P(Lp())P(Lp())PM_{f}P\colon P(\mathrm{L}^{p}(\mathbb{R}))\to P(\mathrm{L}^{p}(\mathbb{R})) of (4.1) (with n=1n=1) identifies to the Toeplitz operator Tf:Hp()Hp()T_{f}\colon\mathrm{H}^{p}(\mathbb{R})\to\mathrm{H}^{p}(\mathbb{R}), gP(fg)g\mapsto P(fg) with symbol ff, where we identify the map PP with its corestriction P|Hp()P|^{\mathrm{H}^{p}(\mathbb{R})}. So the operator TfT_{f} is Fredholm. Using Theorem 8.7, we deduce that

[f],(Lp(),i)K1(C0()),K1(C0(),p)=(4.4)IndexPMfP=IndexTf=windf.\displaystyle\big{\langle}[f],(\mathrm{L}^{p}(\mathbb{R}),\mathrm{i}\mathcal{H})\big{\rangle}_{\mathrm{K}_{1}(\mathrm{C}_{0}(\mathbb{R})),\mathrm{K}^{1}(\mathrm{C}_{0}(\mathbb{R}),\mathscr{L}^{p})}\overset{\eqref{pairing-odd-2}}{=}\operatorname{Index}PM_{f}P=\operatorname{Index}T_{f}=-\operatorname{wind}f.

   

Remark 8.10

Our approach is flexible. We can state variations with a possible weight, matricial versions and UMD\mathrm{UMD}-vector-valued variants.

Remark 8.11

Suppose that p=2p=2. If 0<q<0<q<\infty, It is known that the commutator [F,Mf][F,M_{f}] is in the Schatten class Sq(L2())S^{q}(\mathrm{L}^{2}(\mathbb{R})) if and only if the function ff is in the Besov space Bq,qq()\mathrm{B}^{q}_{q,q}(\mathbb{R}).

Remark 8.12

It should be possible to prove the case p=2p=2 of Theorem 8.7 using Connes character formula for locally compact spectral triples. It suffices to find the result in the literature on the noncommutative measure (using the Dixmier trace) and to check the assumptions of [SuZ23, Theorem 1.2.5].

8.3 The Dirac operator on the space Lp()Lp()\mathrm{L}^{p}(\mathbb{C})\oplus\mathrm{L}^{p}(\mathbb{C}) and the complex Riesz transform

In this section, we will use the classical operators

z¯=def12(x+iy)andz=def12(xiy).\partial_{\overline{z}}\overset{\mathrm{def}}{=}\frac{1}{2}\bigg{(}\frac{\partial}{\partial x}+\mathrm{i}\frac{\partial}{\partial y}\bigg{)}\quad\text{and}\quad\partial_{z}\overset{\mathrm{def}}{=}\frac{1}{2}\bigg{(}\frac{\partial}{\partial x}-\mathrm{i}\frac{\partial}{\partial y}\bigg{)}.

Recall that Δ=4z¯z\Delta=4\partial_{\overline{z}}\partial_{z} The operators z¯\partial_{\overline{z}}, z\partial_{z} and Δ\Delta can be seen as Fourier multipliers on with symbol iπζ-\mathrm{i}\pi\zeta, iπζ¯-\mathrm{i}\pi\overline{\zeta} and |ζ|2|\zeta|^{2}, see [AIM09, p. 99]. In the sequel, we consider the Dirac operator

(8.17) D=def[02z¯2z0].D\overset{\mathrm{def}}{=}\begin{bmatrix}0&-2\partial_{\overline{z}}\\ 2\partial_{z}&0\\ \end{bmatrix}.

acting on a subspace of the Banach space Lp()Lp()\mathrm{L}^{p}(\mathbb{C})\oplus\mathrm{L}^{p}(\mathbb{C}). The square of this operator is given by D2=[4z¯z004zz¯]=[Δ00Δ]D^{2}=\begin{bmatrix}-4\partial_{\overline{z}}\partial_{z}&0\\ 0&-4\partial_{z}\partial_{\overline{z}}\\ \end{bmatrix}=\begin{bmatrix}-\Delta&0\\ 0&-\Delta\\ \end{bmatrix}. In the sequel, we use the notations =def2z\partial\overset{\mathrm{def}}{=}2\partial_{z} and =def2z¯\partial^{*}\overset{\mathrm{def}}{=}-2\partial_{\overline{z}}. We start with a technical result.

Proposition 8.13

Suppose that 1<p<1<p<\infty. The family

(8.18) {t(Idt2Δ)1:t>0}\Big{\{}t\partial(\mathrm{Id}-t^{2}\Delta)^{-1}:t>0\Big{\}}

of operators of B(Lp())\mathrm{B}(\mathrm{L}^{p}(\mathbb{C})) is RR-bounded.

Proof : Note that the Riesz transform (Δ)12:Lp()Lp()\partial(-\Delta)^{-\frac{1}{2}}\colon\mathrm{L}^{p}(\mathbb{C})\to\mathrm{L}^{p}(\mathbb{C}) is a well-defined bounded operator by [AIM09, Corollary 4.5.1 p. 127]. Suppose that t>0t>0. A standard functional calculus argument gives

(8.19) t(Idt2Δ)1=(Δ)12((t2Δ)12(Idt2Δ)1).\displaystyle t\partial(\mathrm{Id}-t^{2}\Delta)^{-1}=\partial(-\Delta)^{-\frac{1}{2}}\Big{(}(-t^{2}\Delta)^{\frac{1}{2}}(\mathrm{Id}-t^{2}\Delta)^{-1}\Big{)}.

By [HvNVW18, Theorem 10.2.25 p. 391], note that the Laplacian Δ-\Delta has a bounded H(Σθ)\mathrm{H}^{\infty}(\Sigma_{\theta}) functional calculus for any angle θ>0\theta>0. Moreover, the Banach space Lp()\mathrm{L}^{p}(\mathbb{C}) is UMD\mathrm{UMD} by [HvNVW16, Proposition 4.2.15 p. 291], hence has the triangular contraction property (Δ)(\Delta) by [HvNVW18, Theorem 7.5.9 p. 137]. We deduce by [HvNVW18, Theorem 10.3.4 (2) p. 401] that the operator Δ-\Delta is RR-sectorial. By [HvNVW18, Example 10.3.5 p. 402] applied with α=12\alpha=\frac{1}{2} and β=1\beta=1, we infer that the set

{(t2Δ)12(Idt2Δ)1:t>0}\big{\{}(-t^{2}\Delta)^{\frac{1}{2}}(\mathrm{Id}-t^{2}\Delta)^{-1}:t>0\big{\}}

of operators of B(Lp())\mathrm{B}(\mathrm{L}^{p}(\mathbb{C})) is RR-bounded. Recalling that a singleton is RR-bounded by [HvNVW18, Example 8.1.7 p. 170], we obtain by composition [HvNVW18, Proposition 8.1.19 (3) p. 178] that the set

{(Δ)12((t2Δ)12(Idt2Δ)1):t>0}\Big{\{}\partial(-\Delta)^{-\frac{1}{2}}\Big{(}(-t^{2}\Delta)^{\frac{1}{2}}(\mathrm{Id}-t^{2}\Delta)^{-1}\Big{)}:t>0\Big{\}}

of operators of B(Lp())\mathrm{B}(\mathrm{L}^{p}(\mathbb{C})) is RR-bounded. Hence with (8.19) we conclude that the subset (8.27) is RR-bounded.    

Theorem 8.14

Suppose that 1<p<1<p<\infty. The unbounded operator DD admits a bounded H(Σθbi)\mathrm{H}^{\infty}(\Sigma_{\theta}^{\mathrm{bi}}) functional calculus on the Banach space Lp()Lp()\mathrm{L}^{p}(\mathbb{C})\oplus\mathrm{L}^{p}(\mathbb{C}) for some 0<θ<π20<\theta<\frac{\pi}{2}.

Proof : We will start by showing that the set {it:t,t0}\{\mathrm{i}t:t\in\mathbb{R},t\not=0\} is contained in the resolvent set of DD. We will do this by showing that IditD\mathrm{Id}-\mathrm{i}tD has a two-sided bounded inverse (IditD)1(\mathrm{Id}-\mathrm{i}tD)^{-1} given by

(8.20) [(Idt2Δ)1it(Idt2Δ)1it(Idt2Δ)1(Idt2Δ)1]\begin{bmatrix}(\mathrm{Id}-t^{2}\Delta)^{-1}&\mathrm{i}t(\mathrm{Id}-t^{2}\Delta)^{-1}\partial^{*}\\ \mathrm{i}t\partial(\mathrm{Id}-t^{2}\Delta)^{-1}&(\mathrm{Id}-t^{2}\Delta)^{-1}\end{bmatrix}

acting on Lp()pLp()\mathrm{L}^{p}(\mathbb{C})\oplus_{p}\mathrm{L}^{p}(\mathbb{C}). By Proposition 8.13 and since the operator Δ-\Delta satisfy the property (2.6) of RR-sectoriality, the four entries are bounded. It only remains to check that this matrix defines a two-sided inverse of IditD\mathrm{Id}-\mathrm{i}tD. We have the following equalities of operators acting on domD\operatorname{dom}D.

[(Idt2Δ)1it(IdLpt2Δ)1it(Idt2Δ)1(Idt2Δ)1](IditD)\displaystyle\begin{bmatrix}(\mathrm{Id}-t^{2}\Delta)^{-1}&\mathrm{i}t(\mathrm{Id}_{\mathrm{L}^{p}}-t^{2}\Delta)^{-1}\partial^{*}\\ \mathrm{i}t(\mathrm{Id}-t^{2}\Delta)^{-1}\partial&(\mathrm{Id}-t^{2}\Delta)^{-1}\end{bmatrix}(\mathrm{Id}-\mathrm{i}tD)
=[(Idt2Δ)1it(Idt2Δ)1it(Idt2Δ)1(Idt2Δ)1][IdititId]\displaystyle=\begin{bmatrix}(\mathrm{Id}-t^{2}\Delta)^{-1}&\mathrm{i}t(\mathrm{Id}-t^{2}\Delta)^{-1}\partial^{*}\\ \mathrm{i}t(\mathrm{Id}-t^{2}\Delta)^{-1}\partial&(\mathrm{Id}-t^{2}\Delta)^{-1}\end{bmatrix}\begin{bmatrix}\mathrm{Id}&-\mathrm{i}t\partial^{*}\\ -\mathrm{i}t\partial&\mathrm{Id}\end{bmatrix}
=[(Idt2Δ)1+t2(Idt2Δ)1it(Idt2Δ)1+it(Idt2Δ)1it(Idt2Δ)1it(Idt2Δ)1t2(Idt2Δ)1+(Idt2Δ)1]\displaystyle=\left[\begin{matrix}(\mathrm{Id}-t^{2}\Delta)^{-1}+t^{2}(\mathrm{Id}-t^{2}\Delta)^{-1}\partial^{*}\partial&-\mathrm{i}t(\mathrm{Id}-t^{2}\Delta)^{-1}\partial^{*}+\mathrm{i}t(\mathrm{Id}-t^{2}\Delta)^{-1}\partial^{*}\\ \mathrm{i}t(\mathrm{Id}-t^{2}\Delta)^{-1}\partial-\mathrm{i}t(\mathrm{Id}-t^{2}\Delta)^{-1}\partial&t^{2}(\mathrm{Id}-t^{2}\Delta)^{-1}\partial\partial^{*}+(\mathrm{Id}-t^{2}\Delta)^{-1}\end{matrix}\right]
=[(Idt2Δ)1+t2(Idt2Δ)1Δ0it(Idt2Δ)1it(Idt2Δ)1(Idt2Δ)1(t2+Id)]=[Id00Id]\displaystyle=\left[\begin{matrix}(\mathrm{Id}-t^{2}\Delta)^{-1}+t^{2}(\mathrm{Id}-t^{2}\Delta)^{-1}\Delta&0\\ \mathrm{i}t(\mathrm{Id}-t^{2}\Delta)^{-1}\partial-\mathrm{i}t\big{(}\mathrm{Id}-t^{2}\Delta\big{)}^{-1}\partial&(\mathrm{Id}-t^{2}\Delta)^{-1}(t^{2}\partial\partial^{*}+\mathrm{Id})\end{matrix}\right]=\begin{bmatrix}\mathrm{Id}&0\\ 0&\mathrm{Id}\end{bmatrix}

and similarly

(IditD)[(Idt2Δ)1it(Idt2Δ)1it(Idt2Δ)1(Idt2Δ)1]\displaystyle(\mathrm{Id}-\mathrm{i}tD)\begin{bmatrix}(\mathrm{Id}-t^{2}\Delta)^{-1}&\mathrm{i}t(\mathrm{Id}-t^{2}\Delta)^{-1}\partial^{*}\\ \mathrm{i}t(\mathrm{Id}-t^{2}\Delta)^{-1}\partial&(\mathrm{Id}-t^{2}\Delta)^{-1}\end{bmatrix}
=[IdititId][(Idt2Δ)1it(Idt2Δ)1it(Idt2Δ)1(Idt2Δ)1]\displaystyle=\begin{bmatrix}\mathrm{Id}&-\mathrm{i}t\partial^{*}\\ -\mathrm{i}t\partial&\mathrm{Id}\end{bmatrix}\begin{bmatrix}(\mathrm{Id}-t^{2}\Delta)^{-1}&\mathrm{i}t(\mathrm{Id}-t^{2}\Delta)^{-1}\partial^{*}\\ \mathrm{i}t(\mathrm{Id}-t^{2}\Delta)^{-1}\partial&(\mathrm{Id}-t^{2}\Delta)^{-1}\end{bmatrix}
=[(Idt2Δ)1+t2(Idt2Δ)1it(Idt2Δ)1it((Idt2Δ)1)it(Idt2Δ)1+it(Idt2Δ)1t2(Idt2Δ)1+(Idt2Δ)1]\displaystyle=\left[\begin{matrix}(\mathrm{Id}-t^{2}\Delta)^{-1}+t^{2}\partial^{*}(\mathrm{Id}-t^{2}\Delta)^{-1}\partial&\mathrm{i}t(\mathrm{Id}-t^{2}\Delta)^{-1}\partial^{*}-\mathrm{i}t\partial^{*}\big{(}(\mathrm{Id}-t^{2}\Delta)^{-1}\big{)}\\ -\mathrm{i}t\partial(\mathrm{Id}-t^{2}\Delta)^{-1}+\mathrm{i}t\partial(\mathrm{Id}-t^{2}\Delta)^{-1}&t^{2}\partial(\mathrm{Id}-t^{2}\Delta)^{-1}\partial^{*}+(\mathrm{Id}-t^{2}\Delta)^{-1}\end{matrix}\right]
=[Id00Id].\displaystyle=\begin{bmatrix}\mathrm{Id}&0\\ 0&\mathrm{Id}\end{bmatrix}.

It remains to show that the set {it(itD)1:t0}={(IditD)1:t0}\{\mathrm{i}t(\mathrm{i}t-D)^{-1}:t\not=0\}=\{(\mathrm{Id}-\mathrm{i}tD)^{-1}:t\not=0\} is RR-bounded. For this, observe that the diagonal entries of (8.20) are RR-bounded by the RR-sectoriality of Δ\Delta. The RR-boundedness of the other entries follows from the RR-gradient bounds of Proposition 8.13. Since a set of operator matrices is RR-bounded precisely when each entry is RR-bounded, we conclude that (2.6) is satisfied, i.e. that the operator DD is RR-bisectorial.    

Remark 8.15

We can introduce the homomorphism π:C0()B(Lp()Lp())\pi\colon\mathrm{C}_{0}^{\infty}(\mathbb{C})\to\mathrm{B}(\mathrm{L}^{p}(\mathbb{C})\oplus\mathrm{L}^{p}(\mathbb{C})), f[Mf00Mf]f\mapsto\begin{bmatrix}M_{f}&0\\ 0&M_{f}\\ \end{bmatrix}. It is easy to check that the commutators [D,π(f)][D,\pi(f)] are well-defined bounded operators.

Note that we have

signD=(8.17)D|D|1=[02z¯2z0][Δ00Δ]12=[02z(Δ)120]\operatorname{\mathrm{sign}}D\overset{\eqref{Dirac-operator-complex}}{=}D|D|^{-1}=\begin{bmatrix}0&-2\partial_{\overline{z}}\\ 2\partial_{z}&0\\ \end{bmatrix}\begin{bmatrix}-\Delta&0\\ 0&-\Delta\\ \end{bmatrix}^{-\frac{1}{2}}=\begin{bmatrix}0&*\\ 2\partial_{z}(-\Delta)^{-\frac{1}{2}}&0\\ \end{bmatrix}

Note that the operator z(Δ)12\partial_{z}(-\Delta)^{-\frac{1}{2}} is the complex Riesz transform R:Lp()Lp()R\colon\mathrm{L}^{p}(\mathbb{C})\to\mathrm{L}^{p}(\mathbb{C}) (or complex Hilbert transform) considered in [AIM09, pp. 102-103]. It is a Fourier multiplier with symbol ζ¯|ζ|\frac{\overline{\zeta}}{|\zeta|}. We have an even Banach Fredholm module (Lp()Lp(),signD,[Id00Id])\bigg{(}\mathrm{L}^{p}(\mathbb{C})\oplus\mathrm{L}^{p}(\mathbb{C}),\operatorname{\mathrm{sign}}D,\begin{bmatrix}-\mathrm{Id}&0\\ 0&\mathrm{Id}\\ \end{bmatrix}\bigg{)}.

Remark 8.16

By [WeO93, Exercise 9.C p. 172], we have K0(C0(2))=\mathrm{K}_{0}(\mathrm{C}_{0}(\mathbb{R}^{2}))=\mathbb{Z} and K1(C0(2))=0\mathrm{K}_{1}(\mathrm{C}_{0}(\mathbb{R}^{2}))=0. The index pairing is related to the Bott projector. We skip the well-known explanation.

Remark 8.17

It is worth noting that by [AIM09, p. 102], the (planar) Beurling-Ahlfors operator 𝒮:Lp()Lp()\mathcal{S}\colon\mathrm{L}^{p}(\mathbb{C})\to\mathrm{L}^{p}(\mathbb{C}) defined by

(8.21) (𝒮f)(z)=def1πp.v.f(w)(wz)2dA(w).(\mathcal{S}f)(z)\overset{\mathrm{def}}{=}-\frac{1}{\pi}\operatorname{p.v.}\int_{\mathbb{C}}\frac{f(w)}{(w-z)^{2}}\mathop{}\mathopen{}\mathrm{d}A(w).

is equal to R2R^{2}, with symbol (ζ¯ζ)2=ζ¯ζ\big{(}\frac{\overline{\zeta}}{\zeta}\big{)}^{2}=\frac{\overline{\zeta}}{\zeta}, see [AIM09, Corollary 4.1.1 p. 102]. This operator has the property that it turns z¯\bar{z}-derivatives into zz-derivatives, i.e. we have 𝒯(z¯f)=zf\mathcal{T}(\partial_{\overline{z}}f)=\partial_{z}f for any function fC0()f\in\mathrm{C}^{\infty}_{0}(\mathbb{C}). Note that (Lp()Lp(),[0𝒮1𝒮0],[Id00Id])\bigg{(}\mathrm{L}^{p}(\mathbb{C})\oplus\mathrm{L}^{p}(\mathbb{C}),\begin{bmatrix}0&\mathcal{S}^{-1}\\ \mathcal{S}&0\\ \end{bmatrix},\begin{bmatrix}-\mathrm{Id}&0\\ 0&\mathrm{Id}\\ \end{bmatrix}\bigg{)} is clearly an even Banach Fredholm module, since by [AIM09, Theorem 4.6.14 p. 145] the commutators [𝒯,Mf][\mathcal{T},M_{f}] are compact on Lp()\mathrm{L}^{p}(\mathbb{C}) for any function ff of VMO()\mathrm{VMO}(\mathbb{C}) and since . By the way, it is worth noting that by [AIM09, Theorem 4.6.13 p. 143] the commutators [𝒯,Mf][\mathcal{T},M_{f}] are bounded on the Banach space Lp()\mathrm{L}^{p}(\mathbb{C}) for any function ff belonging to the space BMO()\mathrm{BMO}(\mathbb{C}).

8.4 The Dirac operator on the space Lp(𝕋θ2)Lp(𝕋θ2)\mathrm{L}^{p}(\mathbb{T}^{2}_{\theta})\oplus\mathrm{L}^{p}(\mathbb{T}^{2}_{\theta})

Quantum tori

We will use standard notations and we refer to the papers [CXY13], [FXZ23], [EcI18], [MSX19] and [XXY18] for more information. Let d2d\geqslant 2. To each d×dd\times d real skew-symmetric matrix θ\theta, one may associate a 2-cocycle σθ:d×d𝕋\sigma_{\theta}\colon\mathbb{Z}^{d}\times\mathbb{Z}^{d}\to\mathbb{T} of the group d\mathbb{Z}^{d} defined by σθ(m,n)=defei2m,θn\sigma_{\theta}(m,n)\overset{\mathrm{def}}{=}\mathrm{e}^{\frac{\mathrm{i}}{2}\langle m,\theta n\rangle} where m,ndm,n\in\mathbb{Z}^{d}. We have σ(m,m)=σ(m,m)\sigma(m,-m)=\sigma(-m,m) for any mdm\in\mathbb{Z}^{d}.

We define the dd-dimensional noncommutative torus L(𝕋θd)\mathrm{L}^{\infty}(\mathbb{T}_{\theta}^{d}) as the twisted group von Neumann algebra VN(d,σθ)\mathrm{VN}(\mathbb{Z}^{d},\sigma_{\theta}). One can provide a concrete realization in the following manner. If (εn)nd(\varepsilon_{n})_{n\in\mathbb{Z}^{d}} is the canonical basis of the Hilbert space d2\ell^{2}_{\mathbb{Z}^{d}} and if mdm\in\mathbb{Z}^{d}, we can consider the bounded operator Um:d2d2U_{m}\colon\ell^{2}_{\mathbb{Z}^{d}}\to\ell^{2}_{\mathbb{Z}^{d}} defined by

(8.22) Um(εn)=defσθ(m,n)εm+n,n.U_{m}(\varepsilon_{n})\overset{\mathrm{def}}{=}\sigma_{\theta}(m,n)\varepsilon_{m+n},\quad n\in\mathbb{Z}.

The dd-dimensional noncommutative torus L(𝕋θd)\mathrm{L}^{\infty}(\mathbb{T}_{\theta}^{d}) is the von Neumann subalgebra of B(d2)\mathrm{B}(\ell^{2}_{\mathbb{Z}^{d}}) generated by the *-algebra 𝒫θ=defspan{Um:md}\mathcal{P}_{\theta}\overset{\mathrm{def}}{=}\mathrm{span}\big{\{}U^{m}\ :\ m\in\mathbb{Z}^{d}\big{\}}. Recall that for any m,ndm,n\in\mathbb{Z}^{d} we have

(8.23) UmUn=σθ(m,n)Um+nand(Um)=σθ(m,m)¯Um.U_{m}U_{n}=\sigma_{\theta}(m,n)U_{m+n}\quad\text{and}\quad\big{(}U_{m}\big{)}^{*}=\overline{\sigma_{\theta}(m,-m)}U_{-m}.

The von Neumann algebra L(𝕋θd)\mathrm{L}^{\infty}(\mathbb{T}_{\theta}^{d}) is finite with normalized trace given by τ(x)=defε0,x(ε0)d2\tau(x)\overset{\mathrm{def}}{=}\langle\varepsilon_{0},x(\varepsilon_{0})\rangle_{\ell^{2}_{\mathbb{Z}^{d}}} where xL(𝕋θd)x\in\mathrm{L}^{\infty}(\mathbb{T}_{\theta}^{d}). In particular, we have τ(Um)=δm=0\tau(U_{m})=\delta_{m=0} for any mdm\in\mathbb{Z}^{d}.

Let Δ\Delta be the unbounded operator acting on L(𝕋θd)\mathrm{L}^{\infty}(\mathbb{T}_{\theta}^{d}) defined on the weak* dense subspace 𝒫θ\mathcal{P}_{\theta} by Δ(Um)=def4π2|m|2Um\Delta(U_{m})\overset{\mathrm{def}}{=}4\pi^{2}|m|^{2}U_{m} where |m|=defm12++md2|m|\overset{\mathrm{def}}{=}m_{1}^{2}+\cdots+m_{d}^{2}. Then this operator is weak* closable and its weak* closure is the opposite of a weak* generator of a symmetric Markovian semigroup (Tt)t0(T_{t})_{t\geqslant 0} of operators acting on L(𝕋θd)\mathrm{L}^{\infty}(\mathbb{T}_{\theta}^{d}), called the noncommutative heat semigroup on the noncommutative torus.

For any j=1,,dj=1,\ldots,d, we may define the partial differentiation operators j\partial_{j} by

j(Un)=def2πinjUn,n=(n1,,nd)d.\partial_{j}(U^{n})\overset{\mathrm{def}}{=}2\pi\mathrm{i}n_{j}U^{n},\quad n=(n_{1},\ldots,n_{d})\in\mathbb{Z}^{d}.

Every partial derivation j\partial_{j} can be viewed a densely defined closed unbounded operator acting on the Hilbert space L2(𝕋θd)\mathrm{L}^{2}(\mathbb{T}^{d}_{\theta}).

The Dirac operator

The Dirac operator DD is defined in terms of γ\gamma matrices in direct analogy to commutative tori. Define N=def2d2N\overset{\mathrm{def}}{=}2^{\lfloor\frac{d}{2}\rfloor} and select N×NN\times N complex selfadjoint matrices {γ1,,γd}\{\gamma_{1},\ldots,\gamma_{d}\} satisfying γjγk+γkγj=2δj,k1\gamma_{j}\gamma_{k}+\gamma_{k}\gamma_{j}=2\delta_{j,k}1. Following [EcI18, (B.6) p. 147] and [GVF01, Definition 12.14 p. 545], we define the unbounded densely defined linear operator

(8.24) D=defij=1dγjjD\overset{\mathrm{def}}{=}-\mathrm{i}\sum_{j=1}^{d}\gamma_{j}\otimes\partial_{j}

acting on the complex Hilbert space N2(L2(𝕋θd))\ell^{2}_{N}(\mathrm{L}^{2}(\mathbb{T}^{d}_{\theta})). The operator DD is selfadjoint. By [GVF01, p. 545], we have

(8.25) D2=IdN2Δ.D^{2}=-\mathrm{Id}_{\ell^{2}_{N}}\otimes\Delta.
Example 8.18

If d=2d=2 then N=2N=2 and by [GVF01, p. 545] we have

(8.26) D=i[01i21+i20].D=-\mathrm{i}\begin{bmatrix}0&\partial_{1}-\mathrm{i}\partial_{2}\\ \partial_{1}+\mathrm{i}\partial_{2}&0\\ \end{bmatrix}.

Boundedness of the functional calculus

We will use the following result which says that some Riesz transforms are bounded. It will be proved in a companion paper [Arh24d].

Theorem 8.19

Suppose that 1<p<1<p<\infty. Consider an integer d2d\geqslant 2 and a real skew-symmetric matrix θMd()\theta\in\mathrm{M}_{d}(\mathbb{R}). For any 1jd1\leqslant j\leqslant d, the linear map jΔp12\partial_{j}\Delta_{p}^{-\frac{1}{2}} is bounded from the subspace RanΔp12\operatorname{Ran}\Delta_{p}^{\frac{1}{2}} into the Banach space Lp(𝕋θd)\mathrm{L}^{p}(\mathbb{T}_{\theta}^{d}).

We introduce the map =def1+i2\partial\overset{\mathrm{def}}{=}\partial_{1}+\mathrm{i}\partial_{2}. First, we prove a technical result, similar to Proposition 8.13.

Proposition 8.20

Suppose that 1<p<1<p<\infty. The family

(8.27) {t(Idt2Δ)1:t>0}\Big{\{}t\partial(\mathrm{Id}-t^{2}\Delta)^{-1}:t>0\Big{\}}

of operators of B(Lp(𝕋θ2))\mathrm{B}(\mathrm{L}^{p}(\mathbb{T}^{2}_{\theta})) is RR-bounded.

Proof : Since we have the equality (Δ)12=1(Δ)12+i2(Δ)12\partial(-\Delta)^{-\frac{1}{2}}=\partial_{1}(-\Delta)^{-\frac{1}{2}}+\mathrm{i}\partial_{2}(-\Delta)^{-\frac{1}{2}}, the Riesz transform (Δ)12:Lp(𝕋θ2)Lp(𝕋θ2)\partial(-\Delta)^{-\frac{1}{2}}\colon\mathrm{L}^{p}(\mathbb{T}^{2}_{\theta})\to\mathrm{L}^{p}(\mathbb{T}^{2}_{\theta}) is a well-defined bounded operator by Theorem 8.19. Suppose that t>0t>0. A standard functional calculus argument gives

(8.28) t(Idt2Δ)1=(Δ)12((t2Δ)12(Idt2Δ)1).\displaystyle t\partial(\mathrm{Id}-t^{2}\Delta)^{-1}=\partial(-\Delta)^{-\frac{1}{2}}\Big{(}(-t^{2}\Delta)^{\frac{1}{2}}(\mathrm{Id}-t^{2}\Delta)^{-1}\Big{)}.

By transference, note that the Laplacian Δ-\Delta has a bounded H(Σθ)\mathrm{H}^{\infty}(\Sigma_{\theta}) functional calculus for any angle θ>0\theta>0. Moreover, the noncommutative Lp\mathrm{L}^{p}-space Lp(𝕋θ2)\mathrm{L}^{p}(\mathbb{T}^{2}_{\theta}) is a UMD\mathrm{UMD} Banach space by [PiX03, Corollary 7.7 p. 1494], hence has the triangular contraction property (Δ)(\Delta) by [HvNVW18, Theorem 7.5.9 p. 137]. We deduce by [HvNVW18, Theorem 10.3.4 (2) p. 401] that the operator Δ-\Delta is RR-sectorial. By [HvNVW18, Example 10.3.5 p. 402] applied with α=12\alpha=\frac{1}{2} and β=1\beta=1, we infer that the set

{(t2Δ)12(Idt2Δ)1:t>0}\big{\{}(-t^{2}\Delta)^{\frac{1}{2}}(\mathrm{Id}-t^{2}\Delta)^{-1}:t>0\big{\}}

of operators of B(Lp(𝕋θ2))\mathrm{B}(\mathrm{L}^{p}(\mathbb{T}^{2}_{\theta})) is RR-bounded. Recalling that a singleton is RR-bounded by [HvNVW18, Example 8.1.7 p. 170], we obtain by composition [HvNVW18, Proposition 8.1.19 (3) p. 178] that the set

{(Δ)12((t2Δ)12(Idt2Δ)1):t>0}\Big{\{}\partial(-\Delta)^{-\frac{1}{2}}\Big{(}(-t^{2}\Delta)^{\frac{1}{2}}(\mathrm{Id}-t^{2}\Delta)^{-1}\Big{)}:t>0\Big{\}}

of operators of B(Lp(𝕋θ2))\mathrm{B}(\mathrm{L}^{p}(\mathbb{T}^{2}_{\theta})) is RR-bounded. Hence with (8.28) we conclude that the subset (8.27) is RR-bounded.    

Proposition 8.21

Suppose that 1<p<1<p<\infty. The operator DD is bisectorial and admits a bounded H(Σθbi)\mathrm{H}^{\infty}(\Sigma_{\theta}^{\mathrm{bi}}\big{)} functional calculus on the Banach space Lp(𝕋θ2)pLp(𝕋θ2)\mathrm{L}^{p}(\mathbb{T}^{2}_{\theta})\oplus_{p}\mathrm{L}^{p}(\mathbb{T}^{2}_{\theta}) for some angle 0<θ<π20<\theta<\frac{\pi}{2}.

Proof : We will start by showing that the set {it:t,t0}\{\mathrm{i}t:t\in\mathbb{R},t\not=0\} is contained in the resolvent set of DD. We will do this by showing that IditD\mathrm{Id}-\mathrm{i}tD has a two-sided bounded inverse (IditD)1(\mathrm{Id}-\mathrm{i}tD)^{-1} given by

(8.29) [(Idt2Δ)1it(Idt2Δ)1it(Idt2Δ)1(Idt2Δ)1]\begin{bmatrix}(\mathrm{Id}-t^{2}\Delta)^{-1}&\mathrm{i}t(\mathrm{Id}-t^{2}\Delta)^{-1}\partial^{*}\\ \mathrm{i}t\partial(\mathrm{Id}-t^{2}\Delta)^{-1}&(\mathrm{Id}-t^{2}\Delta)^{-1}\end{bmatrix}

acting on the Banach space Lp(𝕋θ2)pLp(𝕋θ2)\mathrm{L}^{p}(\mathbb{T}^{2}_{\theta})\oplus_{p}\mathrm{L}^{p}(\mathbb{T}^{2}_{\theta}). By Proposition 8.13 and since the operator Δ-\Delta satisfy the property (2.6) of RR-sectoriality, the four entries are bounded. It only remains to check that this matrix defines a two-sided inverse of IditD\mathrm{Id}-\mathrm{i}tD. We have the following equalities of operators acting on domD\operatorname{dom}D.

[(Idt2Δ)1it(IdLpt2Δ)1it(Idt2Δ)1(Idt2Δ)1](IditD)\displaystyle\begin{bmatrix}(\mathrm{Id}-t^{2}\Delta)^{-1}&\mathrm{i}t(\mathrm{Id}_{\mathrm{L}^{p}}-t^{2}\Delta)^{-1}\partial^{*}\\ \mathrm{i}t(\mathrm{Id}-t^{2}\Delta)^{-1}\partial&(\mathrm{Id}-t^{2}\Delta)^{-1}\end{bmatrix}(\mathrm{Id}-\mathrm{i}tD)
=(8.26)[(Idt2Δ)1it(Idt2Δ)1it(Idt2Δ)1(Idt2Δ)1][IdititId]\displaystyle\overset{\eqref{Dirac-dim-2}}{=}\begin{bmatrix}(\mathrm{Id}-t^{2}\Delta)^{-1}&\mathrm{i}t(\mathrm{Id}-t^{2}\Delta)^{-1}\partial^{*}\\ \mathrm{i}t(\mathrm{Id}-t^{2}\Delta)^{-1}\partial&(\mathrm{Id}-t^{2}\Delta)^{-1}\end{bmatrix}\begin{bmatrix}\mathrm{Id}&-\mathrm{i}t\partial^{*}\\ -\mathrm{i}t\partial&\mathrm{Id}\end{bmatrix}
=[(Idt2Δ)1+t2(Idt2Δ)1it(Idt2Δ)1+it(Idt2Δ)1it(Idt2Δ)1it(Idt2Δ)1t2(Idt2Δ)1+(Idt2Δ)1]\displaystyle=\left[\begin{matrix}(\mathrm{Id}-t^{2}\Delta)^{-1}+t^{2}(\mathrm{Id}-t^{2}\Delta)^{-1}\partial^{*}\partial&-\mathrm{i}t(\mathrm{Id}-t^{2}\Delta)^{-1}\partial^{*}+\mathrm{i}t(\mathrm{Id}-t^{2}\Delta)^{-1}\partial^{*}\\ \mathrm{i}t(\mathrm{Id}-t^{2}\Delta)^{-1}\partial-\mathrm{i}t(\mathrm{Id}-t^{2}\Delta)^{-1}\partial&t^{2}(\mathrm{Id}-t^{2}\Delta)^{-1}\partial\partial^{*}+(\mathrm{Id}-t^{2}\Delta)^{-1}\end{matrix}\right]
=[(Idt2Δ)1+t2(Idt2Δ)1Δ0it(Idt2Δ)1it(Idt2Δ)1(Idt2Δ)1(t2+Id)]=[Id00Id]\displaystyle=\left[\begin{matrix}(\mathrm{Id}-t^{2}\Delta)^{-1}+t^{2}(\mathrm{Id}-t^{2}\Delta)^{-1}\Delta&0\\ \mathrm{i}t(\mathrm{Id}-t^{2}\Delta)^{-1}\partial-\mathrm{i}t\big{(}\mathrm{Id}-t^{2}\Delta\big{)}^{-1}\partial&(\mathrm{Id}-t^{2}\Delta)^{-1}(t^{2}\partial\partial^{*}+\mathrm{Id})\end{matrix}\right]=\begin{bmatrix}\mathrm{Id}&0\\ 0&\mathrm{Id}\end{bmatrix}

and similarly

(IditD)[(Idt2Δ)1it(Idt2Δ)1it(Idt2Δ)1(Idt2Δ)1]=[Id00Id].\displaystyle(\mathrm{Id}-\mathrm{i}tD)\begin{bmatrix}(\mathrm{Id}-t^{2}\Delta)^{-1}&\mathrm{i}t(\mathrm{Id}-t^{2}\Delta)^{-1}\partial^{*}\\ \mathrm{i}t(\mathrm{Id}-t^{2}\Delta)^{-1}\partial&(\mathrm{Id}-t^{2}\Delta)^{-1}\end{bmatrix}=\begin{bmatrix}\mathrm{Id}&0\\ 0&\mathrm{Id}\end{bmatrix}.

It remains to show that the set {it(itD)1:t0}={(IditD)1:t0}\{\mathrm{i}t(\mathrm{i}t-D)^{-1}:t\not=0\}=\{(\mathrm{Id}-\mathrm{i}tD)^{-1}:t\not=0\} is RR-bounded. For this, observe that the diagonal entries of (8.20) are RR-bounded by the RR-sectoriality of Δ\Delta. The RR-boundedness of the other entries follows from the RR-gradient bounds of Proposition 8.13. Since a set of operator matrices is RR-bounded precisely when each entry is RR-bounded, we conclude that (2.6) is satisfied, i.e. that the operator DD is RR-bisectorial.    

For any fL(𝕋θd)f\in\mathrm{L}^{\infty}(\mathbb{T}^{d}_{\theta}), we denote Mf:Lp(𝕋θd)Lp(𝕋θd)M_{f}\colon\mathrm{L}^{p}(\mathbb{T}^{d}_{\theta})\to\mathrm{L}^{p}(\mathbb{T}^{d}_{\theta}), gfgg\mapsto fg the operator of left multiplication on the Banach space Lp(𝕋θd)\mathrm{L}^{p}(\mathbb{T}^{d}_{\theta}). Consequently, we can consider the homomorphism π:L(𝕋θ2)B(Lp(𝕋θ2)Lp(𝕋θ2))\pi\colon\mathrm{L}^{\infty}(\mathbb{T}^{2}_{\theta})\to\mathrm{B}(\mathrm{L}^{p}(\mathbb{T}^{2}_{\theta})\oplus\mathrm{L}^{p}(\mathbb{T}^{2}_{\theta})), fMfMff\mapsto M_{f}\oplus M_{f}.

Theorem 8.22

Suppose that 1<p<1<p<\infty. Consider an integer d2d\geqslant 2 and a real skew-symmetric matrix θMd()\theta\in\mathrm{M}_{d}(\mathbb{R}). The triple (𝒜θ,Lp(𝕋θ2)Lp(𝕋θ2),D)(\mathcal{A}_{\theta},\mathrm{L}^{p}(\mathbb{T}^{2}_{\theta})\oplus\mathrm{L}^{p}(\mathbb{T}^{2}_{\theta}),D) is a compact Banach spectral triple.

Proof : For any function f𝒜θf\in\mathcal{A}_{\theta}, an elementary computation reveals that MfdomDdomDM_{f}\cdot\operatorname{dom}D\subset\operatorname{dom}D and that

(8.30) [D,f]=1i[0MfMf0].[D,f]=\frac{1}{\mathrm{i}}\begin{bmatrix}0&M_{\partial f}\\ M_{\partial f}&0\end{bmatrix}.

So the commutator [D,f][D,f] defines a bounded operator on the Banach space Lp(𝕋θ2)pLp(𝕋θ2)\mathrm{L}^{p}(\mathbb{T}^{2}_{\theta})\oplus_{p}\mathrm{L}^{p}(\mathbb{T}^{2}_{\theta}). It is clear that Δ\Delta has compact resolvent on L2(𝕋θ2)\mathrm{L}^{2}(\mathbb{T}^{2}_{\theta}). By a standard interpolation argument (adapt [Dav89, Section 1.6]), we deduce that Δ\Delta has compact resolvent on Lp(𝕋θ2)\mathrm{L}^{p}(\mathbb{T}^{2}_{\theta}) for any 1<p<1<p<\infty. By (8.25), we conclude that DD has compact resolvent on the space Lp(𝕋θ2)Lp(𝕋θ2)\mathrm{L}^{p}(\mathbb{T}^{2}_{\theta})\oplus\mathrm{L}^{p}(\mathbb{T}^{2}_{\theta}). So Definition 7.4 is satisfied.    

Fredholm module

Using the functional calculus, we can define the operator signD\operatorname{\mathrm{sign}}D acting on the Banach space Lp(𝕋θ2)Lp(𝕋θ2)\mathrm{L}^{p}(\mathbb{T}^{2}_{\theta})\oplus\mathrm{L}^{p}(\mathbb{T}^{2}_{\theta}). This operator ca be seen as ¡¡Riesz-Clifford transform¿¿

sign(D)=j=12γjjΔ12.\operatorname{\mathrm{sign}}(D)=\sum_{j=1}^{2}\gamma_{j}\otimes\partial_{j}\Delta^{-\frac{1}{2}}.

Using Proposition 7.5, we obtain the following consequence of Theorem 8.22.

Corollary 8.23

Suppose that 1<p<1<p<\infty. The triple (𝒜θ,Lp(𝕋θ2)Lp(𝕋θ2),signD)(\mathcal{A}_{\theta},\mathrm{L}^{p}(\mathbb{T}^{2}_{\theta})\oplus\mathrm{L}^{p}(\mathbb{T}^{2}_{\theta}),\operatorname{\mathrm{sign}}D) is a Banach Fredholm module.

Remark 8.24

Suppose that p=2p=2. If fL2(𝕋θd)f\in\mathrm{L}^{2}(\mathbb{T}^{d}_{\theta}) satisfies ¯dfSq\,{\raisebox{-0.56905pt}{$\mathchar 22\relax$}\mkern-12.0mu\mathrm{d}}f\in S^{q}, for some q(0,d]q\in(0,d], then by [MSX19, Corollary 1.5] ff is a constant. Moreover, if ff belongs to the noncommutative homogeneous Sobolev space H˙d1(𝕋θd)\dot{\mathrm{H}}^{1}_{d}(\mathbb{T}^{d}_{\theta}), then by [MSX19, Theorem 1.1] ¯df\,{\raisebox{-0.56905pt}{$\mathchar 22\relax$}\mkern-12.0mu\mathrm{d}}f admits a bounded extension, and the extension belongs to the weak Schatten space Sd,(N2(L2(𝕋θ2)))S^{d,\infty}(\ell^{2}_{N}(\mathrm{L}^{2}(\mathbb{T}^{2}_{\theta}))). In particular, the Fredholm module (𝒜θ,L2(𝕋θ2)L2(𝕋θ2),signD)(\mathcal{A}_{\theta},\mathrm{L}^{2}(\mathbb{T}^{2}_{\theta})\oplus\mathrm{L}^{2}(\mathbb{T}^{2}_{\theta}),\operatorname{\mathrm{sign}}D) is qq-summable for any q>2q>2.

8.5 Perturbed Dirac operators and the Kato square root problem

Let A:nMn()A\colon\mathbb{R}^{n}\to\mathrm{M}_{n}(\mathbb{C}) be a measurable bounded function satisfying for some constant κ>0\kappa>0 the ellipticity condition

ReAxξ,ξnκ|ξ|2\operatorname{Re}\langle A_{x}\xi,\xi\rangle_{\mathbb{C}^{n}}\geqslant\kappa|\xi|^{2}

for almost all xnx\in\mathbb{R}^{n} and all ξn\xi\in\mathbb{C}^{n} . We can see AA as an element of the von Neumann algebra L(n,Mn())\mathrm{L}^{\infty}(\mathbb{R}^{n},\mathrm{M}_{n}(\mathbb{C})). We denote the angle of accretivity by ω=defesssupx,ξ|argAxξ,ξ|\omega\overset{\mathrm{def}}{=}\operatorname*{esssup}_{x,\xi}|\arg\langle A_{x}\xi,\xi\rangle|. Consider the multiplication operator MA:L2(n,n)L2(n,n)M_{A}\colon\mathrm{L}^{2}(\mathbb{R}^{n},\mathbb{C}^{n})\to\mathrm{L}^{2}(\mathbb{R}^{n},\mathbb{C}^{n}), uAuu\mapsto Au and the unbounded operator

(8.31) D=def[0divMA0],D\overset{\mathrm{def}}{=}\begin{bmatrix}0&-\mathrm{div}M_{A}\\ \nabla&0\end{bmatrix},

acting on the complex Hilbert space L2()L2(,n)\mathrm{L}^{2}(\mathbb{R})\oplus\mathrm{L}^{2}(\mathbb{R},\mathbb{C}^{n}), where divMa\mathrm{div}M_{a} denotes the composition divMA\mathrm{div}\circ M_{A}. This operator generalizes the one defined in (1.6). We can see this operator as a deformation of the ¡¡Hodge-Dirac operator¿¿

(8.32) 𝒟=def[0div0].\mathscr{D}\overset{\mathrm{def}}{=}\begin{bmatrix}0&-\mathrm{div}\\ \nabla&0\end{bmatrix}.

By [AKM06, Theorem 3.1 (i) p. ], the operator is bisectorial and admits a bounded H(Σθbi)\mathrm{H}^{\infty}(\Sigma^{\mathrm{bi}}_{\theta}) functional calculus for any θ(ω,π2)\theta\in(\omega,\frac{\pi}{2}). So, we can consider the bounded operator signD:L2(n)L2(n,n)L2(n)L2(n,n)\operatorname{\mathrm{sign}}D\colon\mathrm{L}^{2}(\mathbb{R}^{n})\oplus\mathrm{L}^{2}(\mathbb{R}^{n},\mathbb{C}^{n})\to\mathrm{L}^{2}(\mathbb{R}^{n})\oplus\mathrm{L}^{2}(\mathbb{R}^{n},\mathbb{C}^{n}). As explained in [AKM06] the boundedness of the operator signD\operatorname{\mathrm{sign}}D allows to obtain the Kato square root estimate

(divA)12fL2(n)fL2(n,n2),uW1,2(n).\big{\|}(-\mathrm{div}A\nabla)^{\frac{1}{2}}f\big{\|}_{\mathrm{L}^{2}(\mathbb{R}^{n})}\approx\left\|\nabla f\right\|_{\mathrm{L}^{2}(\mathbb{R}^{n},\ell^{2}_{n})},\quad u\in\mathrm{W}^{1,2}(\mathbb{R}^{n}).

We refer to [HvNVW23, pp. 509-513], [Gra14b, Section 4.7] and [Ouh05, Chapter 8] for more information on this famous estimate.

Note that L2(n,n)=n2(L2(n))\mathrm{L}^{2}(\mathbb{R}^{n},\mathbb{C}^{n})=\ell^{2}_{n}(\mathrm{L}^{2}(\mathbb{R}^{n})). If fL(n)f\in\mathrm{L}^{\infty}(\mathbb{R}^{n}), we define the bounded operator πf:L2(n)L2(n,n)L2(n)L2(n,n)\pi_{f}\colon\mathrm{L}^{2}(\mathbb{R}^{n})\oplus\mathrm{L}^{2}(\mathbb{R}^{n},\mathbb{C}^{n})\to\mathrm{L}^{2}(\mathbb{R}^{n})\oplus\mathrm{L}^{2}(\mathbb{R}^{n},\mathbb{C}^{n}) by

(8.33) πf=def[Mf00M~f],fL(n)\pi_{f}\overset{\mathrm{def}}{=}\begin{bmatrix}\mathrm{M}_{f}&0\\ 0&\tilde{\mathrm{M}}_{f}\\ \end{bmatrix},\quad f\in\mathrm{L}^{\infty}(\mathbb{R}^{n})

where the linear map Mf:L2(n)L2(n)\mathrm{M}_{f}\colon\mathrm{L}^{2}(\mathbb{R}^{n})\to\mathrm{L}^{2}(\mathbb{R}^{n}), gfgg\mapsto fg is the multiplication operator by the function ff and where

M~f=defIdn2Mf:n2(L2(n))n2(L2(n)),(h1,,hn)(fh1,,fhn)\tilde{\mathrm{M}}_{f}\overset{\mathrm{def}}{=}\mathrm{Id}_{\ell^{2}_{n}}\otimes\mathrm{M}_{f}\colon\ell^{2}_{n}(\mathrm{L}^{2}(\mathbb{R}^{n}))\to\ell^{2}_{n}(\mathrm{L}^{2}(\mathbb{R}^{n})),\,(h_{1},\ldots,h_{n})\mapsto(fh_{1},\ldots,fh_{n})

is also a multiplication operator (by the function (f,,f)(f,\ldots,f) of m(L(n))\ell^{\infty}_{m}(\mathrm{L}^{\infty}(\mathbb{R}^{n}))). Using [EnN00, Proposition 4.10 p. 31], it is (really) easy to check that π:L(n)B(L2(n)L2(n,n))\pi\colon\mathrm{L}^{\infty}(\mathbb{R}^{n})\to\mathrm{B}(\mathrm{L}^{2}(\mathbb{R}^{n})\oplus\mathrm{L}^{2}(\mathbb{R}^{n},\mathbb{C}^{n})) is an isometric homomorphism.

Proposition 8.25

The triple (Cc(),L2()L2(,n),D)(\mathrm{C}_{c}^{\infty}(\mathbb{R}),\mathrm{L}^{2}(\mathbb{R})\oplus\mathrm{L}^{2}(\mathbb{R},\mathbb{C}^{n}),D) is a locally compact Banach spectral triple.

Proof : We already said that the operator DD is bisectorial and admits a bounded H(Σθbi)\mathrm{H}^{\infty}(\Sigma^{\mathrm{bi}}_{\theta}) functional calculus. Note that D=[0div0][I00MA]=(8.32)𝒟BD=\begin{bmatrix}0&-\mathrm{div}\\ \nabla&0\end{bmatrix}\begin{bmatrix}\mathrm{I}&0\\ 0&M_{A}\end{bmatrix}\overset{\eqref{D-456}}{=}\mathscr{D}B where B=def[I00MA]B\overset{\mathrm{def}}{=}\begin{bmatrix}\mathrm{I}&0\\ 0&M_{A}\end{bmatrix}. It is easy to see that πfB=Bπf\pi_{f}B=B\pi_{f}. Consequently, we have

[D,πf]=DπfπfD=𝒟Bπfπf𝒟B=𝒟πfBπf𝒟B=(𝒟πfBπf𝒟)B=[𝒟,πf]B.[D,\pi_{f}]=D\pi_{f}-\pi_{f}D=\mathscr{D}B\pi_{f}-\pi_{f}\mathscr{D}B=\mathscr{D}\pi_{f}B-\pi_{f}\mathscr{D}B=(\mathscr{D}\pi_{f}B-\pi_{f}\mathscr{D})B=[\mathscr{D},\pi_{f}]B.

Since it is well-known that the commutator [𝒟,πf][\mathscr{D},\pi_{f}] is bounded, we infer that the commutator [D,πf][D,\pi_{f}] is bounded. We have

πfD1=πf(𝒟B)1=πfB1𝒟1=B1πf𝒟1.\pi_{f}D^{-1}=\pi_{f}(\mathscr{D}B)^{-1}=\pi_{f}B^{-1}\mathscr{D}^{-1}=B^{-1}\pi_{f}\mathscr{D}^{-1}.

Since it is well-known that πf𝒟1\pi_{f}\mathscr{D}^{-1} is compact, we conclude that the operator πfD1:RanD¯X\pi_{f}D^{-1}\colon\overline{\operatorname{Ran}D}\to X is also compact.    

Remark 8.26

It is very easy to introduce variants or generalization of this result. We refer to [AKM06] and [AMN97] for the considered operators.

8.6 Perturbed Dirac operators and the Cauchy singular integral operator

Recall that by Rademacher’s theorem, a function g:g\colon\mathbb{R}\to\mathbb{R} is Lipschitz if and only if gg is differentiable almost everywhere on \mathbb{R} and gL()g^{\prime}\in\mathrm{L}^{\infty}(\mathbb{R}). In this case, the Lipschitz constant

Lipg=defsup{|g(x)g(y)||xy|:x,y,xy}.\operatorname{\mathrm{Lip}}g\overset{\mathrm{def}}{=}\sup\bigg{\{}\frac{|g(x)-g(y)|}{|x-y|}:x,y\in\mathbb{R},x\not=y\bigg{\}}.

is equal to gL()\left\|g^{\prime}\right\|_{\mathrm{L}^{\infty}(\mathbb{R})}. In the sequel, we fix a Lipschitz function g:g\colon\mathbb{R}\to\mathbb{R} and we consider the associated Lipschitz curve in the complex plane, which is the graph γ=def{z=x+ig(x):x}\gamma\overset{\mathrm{def}}{=}\{z=x+\mathrm{i}g(x):x\in\mathbb{R}\}.

Suppose that 1<p<1<p<\infty. Consider the unbounded operator Dγ=defB1iddxD_{\gamma}\overset{\mathrm{def}}{=}-B\frac{1}{\mathrm{i}}\frac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}x} acting on a suitable dense domain of Lp()\mathrm{L}^{p}(\mathbb{R}) where BB is the multiplication operator by the function (1+ig)1(1+\mathrm{i}g^{\prime})^{-1}. By [AKM06, Theorem 3.1 (iii) p.  465] and [AKM06, Consequence 3.2 p. 466], for any angle θ(arctanLipg,π2)\theta\in(\arctan\operatorname{\mathrm{Lip}}g,\frac{\pi}{2}), this operator is bisectorial and admits a bounded H(Σθbi)\mathrm{H}^{\infty}(\Sigma_{\theta}^{\mathrm{bi}}) functional calculus. Note that this result is also proved in [QiL19, Theorem 1.4.1 p. 25] and [McQ91, Theorem 7.1 p. 159] using a different but equivalent formulation.

Consequently, the operator signDγ:Lp()Lp()\operatorname{\mathrm{sign}}D_{\gamma}\colon\mathrm{L}^{p}(\mathbb{R})\to\mathrm{L}^{p}(\mathbb{R}) is bounded. Moreover, according to [AKM06, Consequence 3.2 p. 466], we have

sign(Dγ)=Cγ,\operatorname{\mathrm{sign}}(D_{\gamma})=C_{\gamma},

where CγC_{\gamma} is the ¡¡Cauchy singular integral operator on the real line¿¿ defined by

(8.34) (Cγf)(x)=deflimε0iπy(xε,x+ε)f(y)yx+i(g(y)g(x))(1+ig(y))dy.(C_{\gamma}f)(x)\overset{\mathrm{def}}{=}\lim_{\varepsilon\to 0}\frac{\mathrm{i}}{\pi}\int_{y\in{\mathbb{R}}\setminus(x-\varepsilon,x+\varepsilon)}\frac{f(y)}{y-x+\mathrm{i}(g(y)-g(x))}(1+\mathrm{i}g^{\prime}(y))\mathop{}\mathopen{}\mathrm{d}y.

Here the integral exists for almost all xx\in\mathbb{R} and define an element of the Banach space Lp()\mathrm{L}^{p}(\mathbb{R}). We refer to [McQ91, Theorem 1 p. 142] and [McQ91, (b) p. 157] for a proof of this formula stated in a different language. We direct the reader to the excellent paper [Ver21] for a detailed description of numerous applications of this operator and to [ADM96, Section 8], [Gra14b, Section 4.6] and [Jou83, Chapter 7] for more information.

Remark 8.27

If g=0g=0, then we recover the operator 1iddx\frac{1}{\mathrm{i}}\frac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}x}.

Remark 8.28

Actually, the boundedness of the operator CγC_{\gamma} was first proved by Calderón in [Cal77, Theorem 1, p. 1324] in the case where Lipg\operatorname{\mathrm{Lip}}g is small enough and in the general case in the famous paper [CMM82, Théorème I]. The used definition is

(8.35) (𝒞γu)(z)=deflimε012πiζγB(z,ε)u(ζ)ζzdζ,zγ.(\mathcal{C}_{\gamma}u)(z)\overset{\mathrm{def}}{=}\lim_{\varepsilon\to 0}\frac{1}{2\pi\mathrm{i}}\int\limits_{\zeta\in\gamma\setminus B(z,\varepsilon)}\frac{u(\zeta)}{\zeta-z}\mathop{}\mathopen{}\mathrm{d}\zeta,\qquad z\in\gamma.

Making the bi-Lipschitz change of variables γ\mathbb{R}\to\gamma, yy+ig(y)y\mapsto y+\mathrm{i}g(y) and identifying ff with the function h(y)=defu(y+ig(y))h(y)\overset{\mathrm{def}}{=}u(y+\mathrm{i}g(y)) for xx\in\mathbb{R}, this becomes the Cauchy singular integral operator on the real line, up to a multiplicative constant.

As noted in [Gra14b, p. 289], the Lp\mathrm{L}^{p}-boundedness of the operator Cγ:Lp()Lp()C_{\gamma}\colon\mathrm{L}^{p}(\mathbb{R})\to\mathrm{L}^{p}(\mathbb{R}) is equivalent to that of the related operator 𝒞γ:Lp()Lp()\mathscr{C}_{\gamma}\colon\mathrm{L}^{p}(\mathbb{R})\to\mathrm{L}^{p}(\mathbb{R}) defined by

(8.36) 𝒞γ(f)(x)=deflimε0iπy(xε,x+ε)f(y)yx+i(g(y)g(x))dy,fLp().\mathscr{C}_{\gamma}(f)(x)\overset{\mathrm{def}}{=}\lim_{\varepsilon\to 0}\frac{\mathrm{i}}{\pi}\int_{y\in{\mathbb{R}}\setminus(x-\varepsilon,x+\varepsilon)}\frac{f(y)}{y-x+\mathrm{i}(g(y)-g(x))}\mathop{}\mathopen{}\mathrm{d}y,\quad f\in\mathrm{L}^{p}(\mathbb{R}).

Suppose that 1<p<1<p<\infty. Recall that any multiplication operator Mf:Lp()Lp()M_{f}\colon\mathrm{L}^{p}(\mathbb{R})\to\mathrm{L}^{p}(\mathbb{R}), fgfgfg\mapsto fg by a function ff of L()\mathrm{L}^{\infty}(\mathbb{R}) is bounded, see [EnN00, Proposition 4.10 p. 31]. Consequently, we can consider the homomorphism π:L()B(Lp())\pi\colon\mathrm{L}^{\infty}(\mathbb{R})\to\mathrm{B}(\mathrm{L}^{p}(\mathbb{R})), fMff\mapsto M_{f}.

Consider a function f1<q<Llocq()f\in\cup_{1<q<\infty}\mathrm{L}^{q}_{\textup{loc}}(\mathbb{R}) and suppose that 1<p<1<p<\infty. It is proved in [LNWW20, Theorem 1] that the function ff belongs to the space BMO()\mathrm{BMO}(\mathbb{R}) if and only if the commutator [Mf,𝒞γ][M_{f},\mathscr{C}_{\gamma}] is bounded on the Banach space Lp()\mathrm{L}^{p}(\mathbb{R}). In this case, we have

(8.37) [Mf,𝒞γ]Lp()Lp()fBMO().\left\|[M_{f},\mathscr{C}_{\gamma}]\right\|_{\mathrm{L}^{p}(\mathbb{R})\to\mathrm{L}^{p}(\mathbb{R})}\approx\left\|f\right\|_{\mathrm{BMO}(\mathbb{R})}.

We denote by VMO()\mathrm{VMO}(\mathbb{R}) the space of functions of vanishing mean oscillation, defined to be the BMO(\mathrm{BMO}(\mathbb{R})-closure of the set Cc()\mathrm{C}^{\infty}_{c}(\mathbb{R}) of functions of class C\mathrm{C}^{\infty} with compact support. Furthermore, if fBMO()f\in\mathrm{BMO}(\mathbb{R}) it is proved in [LNWW20, Theorem 2] that ff belongs to VMO()\mathrm{VMO}(\mathbb{R}) if and only if the commutator [Mf,𝒞γ][M_{f},\mathscr{C}_{\gamma}] is a compact operator on the Banach space Lp()\mathrm{L}^{p}(\mathbb{R}). It is obvious that this entails similar results for the operator CγC_{\gamma}.

Proposition 8.29

Then (C0(),L2(),Dγ)(\mathrm{C}_{0}^{\infty}(\mathbb{R}),\mathrm{L}^{2}(\mathbb{R}),D_{\gamma}) is a locally compact Banach spectral triple over the algebra C0()\mathrm{C}_{0}^{\infty}(\mathbb{R}).

Proof : Note that the function 1+ig:1+\mathrm{i}g^{\prime}\colon\mathbb{R}\to\mathbb{R} is bounded above and below. Using the notation D=def1iddxD\overset{\mathrm{def}}{=}\frac{1}{\mathrm{i}}\frac{\mathop{}\mathopen{}\mathrm{d}}{\mathop{}\mathopen{}\mathrm{d}x}, for any function fC0()f\in\mathrm{C}_{0}^{\infty}(\mathbb{R}) we have

[Dγ,πf]=DγπfπfDγ=BDπfπfBD=BDπfBπfD=B(DπfπfD)=B[D,πf].[D_{\gamma},\pi_{f}]=D_{\gamma}\pi_{f}-\pi_{f}D_{\gamma}=BD\pi_{f}-\pi_{f}BD=BD\pi_{f}-B\pi_{f}D=B(D\pi_{f}-\pi_{f}D)=B[D,\pi_{f}].

Since it is well-known that the commutator [D,πf][D,\pi_{f}] is bounded, we infer that the commutator [Dγ,πf][D_{\gamma},\pi_{f}] is bounded. We have

πfDγ1=πf(BD)1=πfD1B1.\pi_{f}D_{\gamma}^{-1}=\pi_{f}(BD)^{-1}=\pi_{f}D^{-1}B^{-1}.

Since it is well-known that πfD1\pi_{f}D^{-1} is compact, we conclude that the operator πfDγ1:RanD¯L2()\pi_{f}D_{\gamma}^{-1}\colon\overline{\operatorname{Ran}D}\to\mathrm{L}^{2}(\mathbb{R}) is also compact.    

Proposition 8.30

Suppose that 1<p<1<p<\infty. Then (Lp(),Cγ)(\mathrm{L}^{p}(\mathbb{R}),C_{\gamma}) is an odd Banach Fredholm module over the algebra C0()\mathrm{C}_{0}(\mathbb{R}).

Proof : If p=2p=2, we could use a possible (easy) locally compact extension of Proposition 7.5. We could also use the following reasoning. Recall that the algebra C0()\mathrm{C}_{0}(\mathbb{R}) of continuous functions that vanish at infinity, since C0()\mathrm{C}_{0}(\mathbb{R}) is a subspace of VMO()\mathrm{VMO}(\mathbb{R}). So if fC0()f\in\mathrm{C}_{0}(\mathbb{R}) by [LNWW20, Theorem 2] each commutator [Mf,𝒞γ][M_{f},\mathscr{C}_{\gamma}] is a compact operator on the Banach space Lp()\mathrm{L}^{p}(\mathbb{R}). So Definition 3.1 is satisfied.    

Remark 8.31

It is possible to make a similar analysis with Clifford-Cauchy singular integrals on Lipschitz surfaces.

Remark 8.32

If AA and BB are some bounded operators on a complex Hilbert space HH, the previous section section and this lead us to consider ¡¡deformed¿¿ triples (𝒜,BDA,H)(\mathcal{A},BDA,H) of a spectral triple (𝒜,D,H)(\mathcal{A},D,H) where the selfadjoint operator DD is replaced by the deformed (not necessarily selfadjoint) operator BDABDA, possibly bisectorial and admitting a bounded H(Σθbi)\mathrm{H}^{\infty}(\Sigma^{\mathrm{bi}}_{\theta}) functional calculus for some angle θ(0,π2)\theta\in(0,\frac{\pi}{2}).

9 Banach Fredholm modules arising from noncommutative Hardy spaces

9.1 Noncommutative Hardy spaces associated to subdiagonal algebras

First, we explain how introduce Hilbert transform associated to subdiagonal algebras of von Neumann algebras. Such algebras were introduced by Arveson in [Arv67] in the case of finite von Neumann algebras and in [Bek15] for the semifinite case. We refer also to [BlL07], [Ji14], [LaX13], [MaW98], [Ran98], [Ran02] and [PiX03] for more information on Hardy spaces associated with subdiagonal algebras of von Neumann algebras. Let \mathcal{M} be a von Neumann algebra endowed with a normal semifinite faithful trace and let E:\mathrm{E}\colon\mathcal{M}\to\mathcal{M} be a trace preserving conditional expectation onto a von Neumann subalgebra 𝒟\mathcal{D} of \mathcal{M}. Following [Bek15, Definition 2.1 p. 1350], a subdiagonal algebra H()\mathrm{H}^{\infty}(\mathcal{M}) with respect to E\mathrm{E} is a weak* closed subalgebra of \mathcal{M} such that

  1. 1.

    H()(H())=𝒟\mathrm{H}^{\infty}(\mathcal{M})\cap(\mathrm{H}^{\infty}(\mathcal{M}))^{*}=\mathcal{D},

  2. 2.

    the trace preserving conditional expectation E:L()L()\mathrm{E}\colon\mathrm{L}^{\infty}(\mathcal{M})\to\mathrm{L}^{\infty}(\mathcal{M}) is multiplicative on H()\mathrm{H}^{\infty}(\mathcal{M}), i.e. we have E(fg)=E(f)E(g)\mathrm{E}(fg)=\mathrm{E}(f)\mathrm{E}(g) for any f,gH()f,g\in\mathrm{H}^{\infty}(\mathcal{M}),

  3. 3.

    H()+(H())\mathrm{H}^{\infty}(\mathcal{M})+(\mathrm{H}^{\infty}(\mathcal{M}))^{*} is weak* dense in L()\mathrm{L}^{\infty}(\mathcal{M}).

Here, we warn the reader that (H())(\mathrm{H}^{\infty}(\mathcal{M}))^{*} is not the dual of H()\mathrm{H}^{\infty}(\mathcal{M}), the family of the adjoints of the elements of H()\mathrm{H}^{\infty}(\mathcal{M}). In this case, 𝒟\mathcal{D} is called the ¡¡diagonal¿¿ of the algebra H()\mathrm{H}^{\infty}(\mathcal{M}). We also define H0()=def{fH():E(f)=0}\mathrm{H}_{0}^{\infty}(\mathcal{M})\overset{\mathrm{def}}{=}\{f\in\mathrm{H}^{\infty}(\mathcal{M}):\mathrm{E}(f)=0\}. Suppose that 1<p<1<p<\infty. The closures of the subspaces H()Lp()\mathrm{H}^{\infty}(\mathcal{M})\cap\mathrm{L}^{p}(\mathcal{M}) and H0()Lp()\mathrm{H}_{0}^{\infty}(\mathcal{M})\cap\mathrm{L}^{p}(\mathcal{M}) in Lp()\mathrm{L}^{p}(\mathcal{M}) will be denoted by Hp()\mathrm{H}^{p}(\mathcal{M}) and H0p()\mathrm{H}^{p}_{0}(\mathcal{M}). Note that E\mathrm{E} extends to a contractive projection E:Lp()Lp()\mathrm{E}\colon\mathrm{L}^{p}(\mathcal{M})\to\mathrm{L}^{p}(\mathcal{M}) onto the subspace Lp(𝒟)\mathrm{L}^{p}(\mathcal{D}) with kernel H0p()\mathrm{H}^{p}_{0}(\mathcal{M}). By [Bek15, Theorem 4.2 p. 1356], we have a canonical topological direct sum decomposition

(9.1) Lp()=(H0p())Lp(𝒟)H0p().\mathrm{L}^{p}(\mathcal{M})=(\mathrm{H}_{0}^{p}(\mathcal{M}))^{*}\oplus\mathrm{L}^{p}(\mathcal{D})\oplus\mathrm{H}_{0}^{p}(\mathcal{M}).

In this setting any fLp()f\in\mathrm{L}^{p}(\mathcal{M}) admits a unique decomposition

f=h+δ+g, with g,hH0p(),δHp(𝒟).f=h^{*}+\delta+g,\quad\text{ with }g,h\in{\mathrm{H}_{0}^{p}(\mathcal{M})},\delta\in\mathrm{H}^{p}(\mathcal{D}).

The Riesz projection P:Lp()Lp()P\colon\mathrm{L}^{p}(\mathcal{M})\to\mathrm{L}^{p}(\mathcal{M}), defined by

(9.2) P(h+δ+g)=δ+g,P(h^{*}+\delta+g)=\delta+g,

is a bounded projection on the subspace Hp()\mathrm{H}^{p}(\mathcal{M}) for any 1<p<1<p<\infty. The Hilbert transform :Lp()Lp()\mathcal{H}\colon\mathrm{L}^{p}(\mathcal{M})\to\mathrm{L}^{p}(\mathcal{M}) associated to the subdiagonal algebra H()\mathrm{H}^{\infty}(\mathcal{M}) is the bounded operator defined by

(9.3) (f)=defi(h+g).\mathcal{H}(f)\overset{\mathrm{def}}{=}-\mathrm{i}(-h^{*}+g).

Toeplitz operators

Given aL()a\in\mathrm{L}^{\infty}(\mathcal{M}), the Toeplitz operator TaT_{a} with symbol aa is defined to be the map

(9.4) Ta:Hp()Hp(),bP+(ab),T_{a}\colon\mathrm{H}^{p}(\mathcal{M})\to\mathrm{H}^{p}(\mathcal{M}),b\mapsto P_{+}(ab),

where P+:Lp()Hp()P_{+}\colon\mathrm{L}^{p}(\mathcal{M})\to\mathrm{H}^{p}(\mathcal{M}) is the corestriction of the Riesz projection P:Lp()Lp()P\colon\mathrm{L}^{p}(\mathcal{M})\to\mathrm{L}^{p}(\mathcal{M}) onto the subspace Hp()\mathrm{H}^{p}(\mathcal{M}).

Banach Fredholm module and pairing with the K\mathrm{K}-theory

Let 𝒜\mathcal{A} be a unital subalgebra of \mathcal{M}. Under suitable assumption, we introduce a Banach Fredholm module and we describe the pairing. We consider the homomorphism π:𝒜B(Lp())\pi\colon\mathcal{A}\to\mathrm{B}(\mathrm{L}^{p}(\mathcal{M})), aMaa\mapsto M_{a}, where Ma:Lp()Lp()M_{a}\colon\mathrm{L}^{p}(\mathcal{M})\to\mathrm{L}^{p}(\mathcal{M}), yayy\mapsto ay is the multiplication operator by aa. We consider the bounded operator

(9.5) F=defi+E:Lp()Lp().F\overset{\mathrm{def}}{=}\mathrm{i}\mathcal{H}+\mathrm{E}\colon\mathrm{L}^{p}(\mathcal{M})\to\mathrm{L}^{p}(\mathcal{M}).
Proposition 9.1

Suppose that 1<p<1<p<\infty. Assume that the commutator [,Ma][\mathcal{H},M_{a}] is a compact operator acting on Lp()\mathrm{L}^{p}(\mathcal{M}) for any a𝒜a\in\mathcal{A} and that the algebra 𝒟\mathcal{D} is finite-dimensional. Then (Lp(),F)(\mathrm{L}^{p}(\mathcal{M}),F) is an odd Banach Fredholm module. For any invertible element a𝒜a\in\mathcal{A}, the Toeplitz operator TaT_{a} is Fredholm and we have

(9.6) [a],(Lp(),F)K1(𝒜),K1(𝒜,ncp)=IndexTa.\big{\langle}[a],(\mathrm{L}^{p}(\mathcal{M}),F)\big{\rangle}_{\mathrm{K}_{1}(\mathcal{A}),\mathrm{K}^{1}(\mathcal{A},\mathscr{L}^{p}_{\mathrm{nc}})}=\operatorname{Index}T_{a}.

Proof : Since the map E:Lp()Lp()\mathrm{E}\colon\mathrm{L}^{p}(\mathcal{M})\to\mathrm{L}^{p}(\mathcal{M}) is a projection on the subspace Lp(𝒟)\mathrm{L}^{p}(\mathcal{D}), we have with (9.3)

F2=(i+E)2=Id.F^{2}=(\mathrm{i}\mathcal{H}+\mathrm{E})^{2}=\mathrm{Id}.

Moreover, for any a𝒜a\in\mathcal{A} we see that

[F,Ma]=[i+E,Ma]=i[,Ma]+[E,Ma][F,M_{a}]=[\mathrm{i}\mathcal{H}+\mathrm{E},M_{a}]=\mathrm{i}[\mathcal{H},M_{a}]+[\mathrm{E},M_{a}]

is a compact operator, since the latter commutator is finite-rank operator. Hence (Lp(),F)(\mathrm{L}^{p}(\mathcal{M}),F) is an odd Banach Fredholm module. Now, observe that the map P=defId+F2:Lp()Lp()P\overset{\mathrm{def}}{=}\frac{\mathrm{Id}+F}{2}\colon\mathrm{L}^{p}(\mathcal{M})\to\mathrm{L}^{p}(\mathcal{M}) is the Riesz projection. Let a𝒜a\in\mathcal{A} be an invertible element. We deduce that PaP:P(Lp())P(Lp())PaP\colon P(\mathrm{L}^{p}(\mathcal{M}))\to P(\mathrm{L}^{p}(\mathcal{M})) is a Fredholm operator. This operator identifies to the Toeplitz operator Ta:Hp()Hp()T_{a}\colon\mathrm{H}^{p}(\mathcal{M})\to\mathrm{H}^{p}(\mathcal{M}) and that

[a],(Lp(),F)K1(𝒜),K1(𝒜,ncp)=(4.4)IndexPaP:P(Lp())P(Lp())=(9.4)Ta.\displaystyle\big{\langle}[a],(\mathrm{L}^{p}(\mathcal{M}),F)\big{\rangle}_{\mathrm{K}_{1}(\mathcal{A}),\mathrm{K}^{1}(\mathcal{A},\mathscr{L}^{p}_{\mathrm{nc}})}\overset{\eqref{pairing-odd-2}}{=}\operatorname{Index}PaP\colon P(\mathrm{L}^{p}(\mathcal{M}))\to P(\mathrm{L}^{p}(\mathcal{M}))\overset{\eqref{Toeplitz}}{=}T_{a}.

   

9.2 Classical Hardy spaces

The fundamental example of subdiagonal algebras is the algebra H(𝕋)\mathrm{H}^{\infty}(\mathbb{T}) on the unit circle. In this case the index pairing is described in Proposition 8.3. We can also state a similar result for the matrix-valued Hardy space H(𝕋,Mn)\mathrm{H}^{\infty}(\mathbb{T},\mathrm{M}_{n}), which be of interest for the multivariate prediction theory. This algebra is a subalgebra of the von Neumann algebra L(𝕋,Mn)\mathrm{L}^{\infty}(\mathbb{T},\mathrm{M}_{n}).

9.3 Fredholm module associated to the discrete Schur-Hilbert transform on SnpS^{p}_{n}

Consider the trace preserving conditional expectation E:MnMn\mathrm{E}\colon\mathrm{M}_{n}\to\mathrm{M}_{n} onto the diagonal subalgebra 𝒟=n\mathcal{D}=\ell^{\infty}_{n} of Mn\mathrm{M}_{n}, where Mn\mathrm{M}_{n} is equipped with its canonical trace. The ¡¡upper triangle subalgebra¿¿ Tn\mathrm{T}_{n} of all upper triangular matrices is a finite subdiagonal algebra of Mn\mathrm{M}_{n}. Suppose that 1<p<1<p<\infty. In this context, the Hilbert transform :SnpSnp\mathcal{H}\colon S^{p}_{n}\to S^{p}_{n} is the ¡¡discrete Schur-Hilbert transform¿¿ defined as being the Schur multiplier with symbol [isign(ij)][-\mathrm{i}\,\operatorname{\mathrm{sign}}(i-j)]. The Riesz projection P:SnpSnpP\colon S^{p}_{n}\to S^{p}_{n} is the triangular projection. We consider the representation π:MnB(Snp)\pi\colon\mathrm{M}_{n}\to\mathrm{B}(S^{p}_{n}), aMaa\mapsto M_{a} of the matrix algebra Mn\mathrm{M}_{n}, where Ma:SnpSnpM_{a}\colon S^{p}_{n}\to S^{p}_{n}, yayy\mapsto ay is the multiplication operator by the matrix aa. The couple (Snp,F)(S^{p}_{n},F) is obviously an odd Banach Fredholm module. The assumptions of Proposition 9.1 are obviously satisfied for the algebra 𝒜=Mn\mathcal{A}=\mathrm{M}_{n}. Since K1(Mn)=0\mathrm{K}_{1}(\mathrm{M}_{n})=0, we obtain the following result.

Proposition 9.2

Suppose that 1<p<1<p<\infty. For any invertible matrix aMna\in\mathrm{M}_{n}, we have

(9.7) [a],(Snp,F)K1(Mn),K1(Mn,ncp)=0.\big{\langle}[a],(S^{p}_{n},F)\big{\rangle}_{\mathrm{K}_{1}(\mathrm{M}_{n}),\mathrm{K}^{1}(\mathrm{M}_{n},\mathscr{L}^{p}_{\mathrm{nc}})}=0.
Remark 9.3

It is possible to state in the same spirit more complicated statements for nest algebras.

9.4 Banach Fredholm modules on reduced group C\mathrm{C}^{*}-algebras

Orderable groups

An order relation \preceq on a group GG is left-invariant (resp. right-invariant) if for all s,tGs,t\in G such that sts\preceq t, one has rsrtrs\preceq rt (resp. srtrsr\preceq tr) for all rGr\in G. The relation is bi-invariant if it is simultaneously left-invariant and right-invariant. We will use the term left-ordering (resp. right-ordering, bi-ordering) for referring to a left-invariant (resp. right-invariant, right-invariant and left-invariant) total order on a group.

Following [ClR16, p. 2], we will say that a group GG is left-orderable (resp. right-orderable, bi-orderable) if it admits a total order which is invariant by the left (resp. by the right, simultaneously by the left and right). We also refer to [DNR16] for more information on these classes of groups. We will write rsr\prec s when rsr\preceq s with rsr\neq s. Note that by [KaT08, Proposition 7.5 p. 277] or [ClR16, Proposition 1.3 p. 2], any orderable group GG is torsion-free.

It is known [DNR16, Proposition 1.1.8 p. 11] [ClR16, Theorem 2.23 p. 28] that a countable group is left-orderable if and only if it is isomorphic to a subgroup of the set of orientation-preserving homeomorphisms of \mathbb{R}.

Example 9.4

Following [ClR16, p. 14], we say that a group is indicable if it has the group of integers \mathbb{Z} as a quotient, and locally indicable if each of its nontrivial finitely-generated sub-groups is indicable. Locally indicable groups are left-orderable by [ClR16, Corollary 1.51 p. 14].

Very large classes of groups are left-orderable as the following list shows.

Example 9.5

Every free group 𝔽n\mathbb{F}_{n} with an arbitrary number nn (possibly infinite) of generators is bi-orderable, as shown in [ClR16, Theorem 3.4 p. 35] or [KaT08, Corollary 7.12 p. 280].

Example 9.6

If an orientable, irreducible2221. A 3-manifold MM is irreducible if every smooth 2-sphere in MM bounds a 3-dimensional ball in MM. 3-manifold MM has infinite first homology group, then by [ClR16, Theorem 5.3 p. 66] its fundamental group π1(M)\pi_{1}(M) is locally indicable, hence left-orderable.

Example 9.7

If KK is a knot, then the fundamental group π1(S3\K)\pi_{1}(S^{3}\backslash K) of its complement S3\KS^{3}\backslash K is called the knot group of KK. By [ClR16, Theorem 4.9 p. 49], such a group is locally indicable, and consequently left-orderable.

Example 9.8

Braid groups

(9.8) Bn=defσ1,,σn1|σiσj=σjσi,|ij|2σiσjσi=σjσiσj,|ij|=1,\mathrm{B}_{n}\overset{\mathrm{def}}{=}\left\langle\sigma_{1},\ldots,\sigma_{n-1}\left|\begin{array}[]{cl}\sigma_{i}\sigma_{j}=\sigma_{j}\sigma_{i},&|i-j|\geqslant 2\\ \sigma_{i}\sigma_{j}\sigma_{i}=\sigma_{j}\sigma_{i}\sigma_{j},&|i-j|=1\end{array}\right.\right\rangle,

where n1n\geqslant 1 is an integer, are left-orderable, according to [KaT08, Theorem 7.15 p. 283], or [ClR16, Theorem 7.23 p. 103]. Note that B1={0}\mathrm{B}_{1}=\{0\} and B2=\mathrm{B}_{2}=\mathbb{Z}. It is worth noting that the group B3\mathrm{B}_{3} is isomorphic to the knot group of the trefoil knot by [ClR16, p. 93]. In particular, it is an infinite non-abelian group. For any integer n3n\geqslant 3, the group Bn\mathrm{B}_{n} is not bi-orderable according to [DDRW08, Proposition 1.2 p. 12].

Example 9.9

The infinite braid group

(9.9) B=defσ1,σ2,|σiσj=σjσi,|ij|2σiσjσi=σjσiσj,|ij|=1,\mathrm{B}_{\infty}\overset{\mathrm{def}}{=}\left\langle\sigma_{1},\sigma_{2},\ldots\left|\begin{array}[]{cl}\sigma_{i}\sigma_{j}=\sigma_{j}\sigma_{i},&|i-j|\geqslant 2\\ \sigma_{i}\sigma_{j}\sigma_{i}=\sigma_{j}\sigma_{i}\sigma_{j},&|i-j|=1\end{array}\right.\right\rangle,

is left-orderable by [ClR16, Problem 10.11 p. 136].

Example 9.10

According to [ClR16, Theorem 7.10 p. 95], the pure braid group Pn\mathrm{P}_{n} is bi-orderable for any integer n2n\geqslant 2.

Example 9.11

Torsion-free nilpotent groups are bi-orderable, as shown in [DNR16, p. 15]. By the way, we refer to the paper [Sud17] for some information on the K-theory of the reduced group C\mathrm{C}^{*}-algebra of these groups.

Example 9.12

For each integer n2n\geqslant 2, the Baumslag-Solitar group BS(1,n)=defa,ba1ba=bn\mathrm{BS}(1,n)\overset{\mathrm{def}}{=}\langle a,b\mid a^{-1}ba=b^{n}\rangle is bi-orderable by [DNR16, p. 17]. The K-theory groups of its reduced group C\mathrm{C}^{*}-algebra are described in [PoV18]. Actually, we have K0(C(BS(1,n)))=\mathrm{K}_{0}(\mathrm{C}^{*}(\mathrm{BS}(1,n)))=\mathbb{Z} and K1(C(BS(1,n)))=/(n1)\mathrm{K}_{1}(\mathrm{C}^{*}(\mathrm{BS}(1,n)))=\mathbb{Z}\oplus\mathbb{Z}/(n-1)\mathbb{Z}.

Example 9.13

The Thompson group F\mathrm{F} is bi-orderable. See Example LABEL:Example-Thompson below for more information.

Example 9.14

If MM is an orientable closed Seifert fibered manifold with infinite first homology group, then the fundamental group π1(M)\pi_{1}(M) is left-orderable, according to [ClR16, Theorem 6.12 p. 83].

Example 9.15

The fundamental groups of all closed surfaces, orientable or not, with the exceptions of the projective plane and the Klein bottle are bi-orderable by [RoW01, Theorem 3 p. 314] or [ClR16, Theorem 3.11 p. 36]. If 𝒦\mathscr{K} is the Klein Bottle then by [ClR16, Problem 1.10 p. 4] its fundamental group π1(𝒦)=x,yxyx1=y1\pi_{1}(\mathscr{K})=\langle x,y\mid xyx^{-1}=y^{-1}\rangle is not bi-orderable. However, it is left-orderable by [RoW01, Theorem 3 p. 314]. Finally, the fundamental group π1(P2())=/2\pi_{1}(\mathrm{P}_{2}(\mathbb{R}))=\mathbb{Z}/2\mathbb{Z} of the projective plane P2()\mathrm{P}_{2}(\mathbb{R}) is not left-orderable (see also [ClR16, Proposition 6.13 p. 84]). A description of the K-theory groups of the reduced group C\mathrm{C}^{*}-algebras of these ¡¡surface groups¿¿ is given in [BBV99].

Example 9.16

Let Σg,nb\Sigma^{b}_{g,n} be an orientable connected surface with genus gg, bb boundary components, and nn punctures3332. Recall that P={p1,,pn}P=\{p_{1},\ldots,p_{n}\} is a finite subset of distinct points ¡¡removed¿¿ from the surface. p1,,pnp_{1},\ldots,p_{n}. If b>0b>0 and if the Euler characteristic of Σg,nb\Sigma^{b}_{g,n} satisfies χ(Σg,nb)<0\chi(\Sigma^{b}_{g,n})<0 then by [ClR16, Theorem 7.29 p. 107], the mapping class group Mod(Σg,nb)\mathrm{Mod}(\Sigma^{b}_{g,n}) is left-orderable. This group is the group of isotopy classes of orientation-preserving homeomorphisms that restrict to the identity on the boundary of Σg,nb\Sigma^{b}_{g,n}.

Example 9.17

Let MM be an nn-dimensional connected piecewise linear manifold (PL manifold). Consider a non-empty closed (n1)(n-1)-dimensional PL submanifold KK of MM. Then by [ClR16, Theorem 8.12 p. 119] the group PL+(M,K)\mathrm{PL}_{+}(M,K) of orientation-preserving PL homeomorphisms of MM, fixed on KK, is locally indicable and therefore left-orderable.

Example 9.18

Let MM be a connected smooth nn-manifold and let KK be a non-empty closed (n1)(n-1)-submanifold of MM. According to [ClR16, Theorem 8.13 p. 119] the group Diff+1(M,K)\mathrm{Diff}^{1}_{+}(M,K) of orientation-preserving homeomorphisms of MM which are fixed on KK and which are continuously differentiable is locally indicable, hence left-orderable.

Example 9.19

The universal cover PSL~2()\widetilde{\mathrm{PSL}}_{2}(\mathbb{R}) of the group PSL2()\mathrm{PSL}_{2}(\mathbb{R}) is left-orderable by [ClR16, Example 1.15 p. 5].

Finally, note that both left-orderability and bi-orderability are clearly preserved under taking subgroups and finite products using lexicographic ordering, see [ClR16, Example 1.6 p. 3]

Fourier multipliers

Let GG be a discrete group. We denote by VN(G)\mathrm{VN}(G) its group von Neumann algebra. Suppose that 1p<1\leqslant p<\infty. We say that a complex function φ:G\varphi\colon G\to\mathbb{C} induces a bounded Fourier multiplier Mφ:Lp(VN(G))Lp(VN(G))\mathrm{M}_{\varphi}\colon\mathrm{L}^{p}(\mathrm{VN}(G))\to\mathrm{L}^{p}(\mathrm{VN}(G)) if the linear map [G][G]\mathbb{C}[G]\to\mathbb{C}[G], λsφsλs\lambda_{s}\mapsto\varphi_{s}\lambda_{s} extends to a bounded map MφM_{\varphi} on the noncommutative Lp\mathrm{L}^{p}-space Lp(VN(G))\mathrm{L}^{p}(\mathrm{VN}(G)), we use the canonical normalized normal finite faithful trace τ\tau on the algebra VN(G)\mathrm{VN}(G) for defining the noncommutative Lp\mathrm{L}^{p}-space. Here [G]\mathbb{C}[G] is the group algebra of GG. We say that φ\varphi is the symbol of Mφ\mathrm{M}_{\varphi}.

Hilbert transforms

Assume that GG is a left-orderable discrete group and endowed with a left-order \preceq. In this context, we can introduce the function sign:G\operatorname{\mathrm{sign}}\colon G\to\mathbb{C} defined by

signs=def{1 if es0 if s=e1 if se.\operatorname{\mathrm{sign}}s\overset{\mathrm{def}}{=}\begin{cases}1&\text{ if }e\prec s\\ 0&\text{ if }s=e\\ -1&\text{ if }s\prec e\end{cases}.

Moreover, we can consider the subalgebra 𝒟=def1\mathcal{D}\overset{\mathrm{def}}{=}\mathbb{C}1 of the group von Neumann algebra VN(G)\mathrm{VN}(G). The linear map E:VN(G)VN(G)\mathrm{E}\colon\mathrm{VN}(G)\to\mathrm{VN}(G), fτ(f)1f\mapsto\tau(f)1 is a trace preserving normal faithful conditional expectation onto the ¡¡diagonal¿¿ subalgebra 𝒟\mathcal{D}. If GG is left-orderable and endowed with a left-order \preceq then it is essentially4443. The proposed definition in H(VN(G))\mathrm{H}^{\infty}(\mathrm{VN}(G)) is slightly incorrect since the ¡¡noncommutative Fourier series¿¿ of elements in VN(G)\mathrm{VN}(G) does not converges in general. observed in [GPX22] that

H(VN(G))=def{finite sums esasλs:as}¯w\mathrm{H}^{\infty}(\mathrm{VN}(G))\overset{\mathrm{def}}{=}\overline{\bigg{\{}\text{finite sums }\sum_{e\preceq s}a_{s}\,\lambda_{s}:a_{s}\in\mathbb{C}\bigg{\}}}^{\mathrm{w}*}

is a subdiagonal algebra whose Hilbert transform :Lp(VN(G))Lp(VN(G))\mathcal{H}\colon\mathrm{L}^{p}(\mathrm{VN}(G))\to\mathrm{L}^{p}(\mathrm{VN}(G)) coincides with the bounded Fourier multiplier with symbol isign-\mathrm{i}\operatorname{\mathrm{sign}}. Actually the boundedness of the last operator was already known for abelian groups [ABG90].

The map F=(9.5)i+E:Lp(VN(G))Lp(VN(G))F\overset{\eqref{def-de-F}}{=}\mathrm{i}\mathcal{H}+\mathrm{E}\colon\mathrm{L}^{p}(\mathrm{VN}(G))\to\mathrm{L}^{p}(\mathrm{VN}(G)) is the Fourier multiplier defined by the symbol φ:G\varphi\colon G\to\mathbb{C} defined by

φs=def{1 if es1 if se.\varphi_{s}\overset{\mathrm{def}}{=}\begin{cases}1&\text{ if }e\preceq s\\ -1&\text{ if }s\prec e\end{cases}.

Finally, the linear map P+:Lp()Hp()P_{+}\colon\mathrm{L}^{p}(\mathcal{M})\to\mathrm{H}^{p}(\mathcal{M}) satisfies P+(λs)=λs=0P_{+}(\lambda_{s})=\lambda_{s}=0 for any sGs\in G satisfying ese\preceq s and P+(λs)=0P_{+}(\lambda_{s})=0 otherwise.

The following result says that the assumptions of Proposition 9.1 are satisfied.

Proposition 9.20

Suppose that 1<p<1<p<\infty. Let GG be a left-orderable discrete group endowed with a left-order. The commutator [,Ma][\mathcal{H},M_{a}] is a compact operator acting on the Banach space Lp(VN(G))\mathrm{L}^{p}(\mathrm{VN}(G)) for any element aa of the reduced group C\mathrm{C}^{*}-algebra Cred(G)\mathrm{C}^{*}_{\mathrm{red}}(G) of the group GG. Moreover, (Lp(VN(G)),F)(\mathrm{L}^{p}(\mathrm{VN}(G)),F) is an odd Banach Fredholm module.

Proof : For any s,tGs,t\in G, we have

[,Mλs](λt)=Mλs(λt)Mλs(λt)=(λst)+isign(t)Mλsλt\displaystyle\left[\mathcal{H},M_{\lambda_{s}}\right](\lambda_{t})=\mathcal{H}M_{\lambda_{s}}(\lambda_{t})-M_{\lambda_{s}}\mathcal{H}(\lambda_{t})=\mathcal{H}(\lambda_{st})+\mathrm{i}\operatorname{\mathrm{sign}}(t)M_{\lambda_{s}}\lambda_{t}
=isign(st)λst+isign(t)λst=i[sign(st)+sign(t)]λst.\displaystyle=-\mathrm{i}\operatorname{\mathrm{sign}}(st)\lambda_{st}+\mathrm{i}\operatorname{\mathrm{sign}}(t)\lambda_{st}=\mathrm{i}\big{[}-\operatorname{\mathrm{sign}}(st)+\operatorname{\mathrm{sign}}(t)\big{]}\lambda_{st}.

If f=asλsf=\sum a_{s}\lambda_{s} is a finite sum belonging to the group algebra [G]\mathbb{C}[G] then we deduce by linearity that the commutator [,Mf]\left[\mathcal{H},M_{f}\right] is a finite-rank operator. We conclude that for any aCred(G)a\in\mathrm{C}^{*}_{\mathrm{red}}(G) the commutator [,Ma]\left[\mathcal{H},M_{a}\right] is compact, as a limit of finite-rank operators, since the homomorphism π:Cred(G)B(Lp(VN(G)))\pi\colon\mathrm{C}^{*}_{\mathrm{red}}(G)\to\mathrm{B}(\mathrm{L}^{p}(\mathrm{VN}(G))), aMaa\mapsto M_{a} is continuous.    

Remark 9.21

Suppose that p=2p=2. For any sGs\in G, it is easy to check that the singular values of the operator [,Mλs]\left[\mathcal{H},M_{\lambda_{s}}\right] are given by

sn([,Mλs])=|sign(tn)sign(stn)|,s_{n}(\left[\mathcal{H},M_{\lambda_{s}}\right])=|\operatorname{\mathrm{sign}}(t_{n})-\operatorname{\mathrm{sign}}(st_{n})|,

where (tn)(t_{n}) is an enumeration of the elements of GG. So the prevous Fredholm module will rarely be finitely summable.

9.5 The index pairing for the reduced group C\mathrm{C}^{*}-algebra of the braid group B3\mathrm{B}_{3}

[KaT08, Theorem 7.15 p. 283], or [ClR16, Theorem 7.23 p. 103] [DDRW08]

Let n2n\geqslant 2 be an integer. The Dehornoy ordering (also known as the σ\sigma-ordering) of the braid group Bn\mathrm{B}_{n} is a left-ordering introduced in [Deh94] that is defined in terms of representative words of braids as follows. Following [ClR16, p. 95](see also [DDRW08, Definition 1.6 p. 12]), if 1in11\leqslant i\leqslant n-1, a word ww in the generators σ1,,σn1\sigma_{1},\ldots,\sigma_{n-1} is called ii-positive if ww contains at least one occurrence of σi\sigma_{i}, no occurrences of σ1,,σi1\sigma_{1},\ldots,\sigma_{i-1}, and every occurrence of σi\sigma_{i} has positive exponent. Replacing ¡¡positive¿¿ by ¡¡negative¿¿, we obtain the definition of a ii-negative word.

A braid βBn\beta\in\mathrm{B}_{n} is called ii-positive (respectively ii-negative) if it admits a representative word ww in the generators σ1,,σn1\sigma_{1},\ldots,\sigma_{n-1} that is ii-positive (respectively ii-negative). The Dehornoy ordering of the braid group Bn\mathrm{B}_{n} is the ordering whose positive elements are the braids βBn\beta\in\mathrm{B}_{n} that are ii-positive for some ii.

Recall that the non-abelian group B3\mathrm{B}_{3} is defined by

(9.10) B3=(9.9)σ1,σ2σ1σ2σ1=σ2σ1σ2.\mathrm{B}_{3}\overset{\eqref{Braid-groups-presentation}}{=}\left\langle\sigma_{1},\sigma_{2}\mid\sigma_{1}\sigma_{2}\sigma_{1}=\sigma_{2}\sigma_{1}\sigma_{2}\right\rangle.

The 2-positive elements are precisely the elements σ2k\sigma_{2}^{k} with k0k\geqslant 0. Examples of 1-positive element are given by σ2σ1kσ22\sigma_{2}\sigma_{1}^{k}\sigma_{2}^{-2} or σ1k\sigma_{1}^{k} with k0k\geqslant 0.

The K-theory groups of the braid group B3\mathrm{B}_{3} are known. Actually, by [ABARW22, Proposition 6.4 p. 1288], we have isomorphisms

K0(Cred(B3))=K1(Cred(B3))=.\mathrm{K}_{0}(\mathrm{C}^{*}_{\mathrm{red}}(\mathrm{B}_{3}))=\mathrm{K}_{1}(\mathrm{C}^{*}_{\mathrm{red}}(\mathrm{B}_{3}))=\mathbb{Z}.

Moreover, the group K0(Cred(B3))\mathrm{K}_{0}(\mathrm{C}^{*}_{\mathrm{red}}(\mathrm{B}_{3})) is generated by the unit of Cred(B3)\mathrm{C}^{*}_{\mathrm{red}}(\mathrm{B}_{3}) and the group K1(Cred(B3))\mathrm{K}_{1}(\mathrm{C}^{*}_{\mathrm{red}}(\mathrm{B}_{3})) is generated by [λσ1]=[λσ2][\lambda_{\sigma_{1}}]=[\lambda_{\sigma_{2}}]. In the next result, we compute the index pairing of (9.6).

Proposition 9.22

Suppose that 1<p<1<p<\infty. For any integer kk\in\mathbb{Z}, we have

(9.11) [λσ1k],(Lp(VN(B3)),F)K1(Cred(B3)),K1(Cred(B3)),ncp)=k.\big{\langle}[\lambda_{\sigma_{1}^{k}}],(\mathrm{L}^{p}(\mathrm{VN}(\mathrm{B}_{3})),F)\big{\rangle}_{\mathrm{K}_{1}(\mathrm{C}^{*}_{\mathrm{red}}(\mathrm{B}_{3})),\mathrm{K}^{1}(\mathrm{C}^{*}_{\mathrm{red}}(\mathrm{B}_{3})),\mathscr{L}^{p}_{\mathrm{nc}})}=-k.

Proof : If i=1i=1 or i=2i=2, we have for any sB3s\in\mathrm{B}_{3}

Tλσik(λs)=P+(λσikλs)=P+(λσiks).T_{\lambda_{\sigma_{i}^{k}}}(\lambda_{s})=P_{+}(\lambda_{\sigma_{i}^{k}}\lambda_{s})=P_{+}(\lambda_{\sigma_{i}^{k}s}).

Suppose that k0k\geqslant 0. If ss is 1-positive then λσ1ks\lambda_{\sigma_{1}^{k}s} is also 1-positive, so Tλσ1k(λs)=λσ1ksT_{\lambda_{\sigma_{1}^{k}}}(\lambda_{s})=\lambda_{\sigma_{1}^{k}s}. If ss is 2-positive then λσ2ks\lambda_{\sigma_{2}^{k}s} is also 2-positive. Hence Tλσ2k(λs)=λσ2ksT_{\lambda_{\sigma_{2}^{k}}}(\lambda_{s})=\lambda_{\sigma_{2}^{k}s}. If s=es=e then Tλσ2k(e)=λσ2kT_{\lambda_{\sigma_{2}^{k}}}(e)=\lambda_{\sigma_{2}^{k}}. By linearity, it is easy to check that KerTλσ1k={0}\operatorname{Ker}T_{\lambda_{\sigma_{1}^{k}}}=\{0\}. Consequently, we have

IndexTλσ1k=(2.1)dimKerTλσ1kdimLp(VN(B3))/RanTλσ1k\displaystyle\operatorname{Index}T_{\lambda_{\sigma_{1}^{k}}}\overset{\eqref{Fredholm-Index}}{=}\dim\operatorname{Ker}T_{\lambda_{\sigma_{1}^{k}}}-\dim\mathrm{L}^{p}(\mathrm{VN}(\mathrm{B}_{3}))/\operatorname{Ran}T_{\lambda_{\sigma_{1}^{k}}}
=dimLp(VN(B3))/RanTλσ1k=k.\displaystyle=-\dim\mathrm{L}^{p}(\mathrm{VN}(\mathrm{B}_{3}))/\operatorname{Ran}T_{\lambda_{\sigma_{1}^{k}}}=-k.

We conclude with (9.6).

Now, suppose that k<0k<0. If s=σ2rs=\sigma_{2}^{r} is 22-positive then σ2ks=σ2kσ2r=σ2r+k\sigma_{2}^{k}s=\sigma_{2}^{k}\sigma_{2}^{r}=\sigma_{2}^{r+k} is 2-positive if and only if r+k0r+k\geqslant 0. It is obvious that the left multiplication by σ2k\sigma_{2}^{k} transforms any 1-positive element into a 1-positive element. So, it is easy to check that dimKerTλσ2k=k\dim\operatorname{Ker}T_{\lambda_{\sigma_{2}^{k}}}=k. We conclude that

IndexTλσ1k=(2.1)dimKerTλσ1kdimLp(VN(B3))/RanTλσ1k=dimKerTλσ1k=k.\displaystyle\operatorname{Index}T_{\lambda_{\sigma_{1}^{k}}}\overset{\eqref{Fredholm-Index}}{=}\dim\operatorname{Ker}T_{\lambda_{\sigma_{1}^{k}}}-\dim\mathrm{L}^{p}(\mathrm{VN}(\mathrm{B}_{3}))/\operatorname{Ran}T_{\lambda_{\sigma_{1}^{k}}}=\dim\operatorname{Ker}T_{\lambda_{\sigma_{1}^{k}}}=k.

   

Remark 9.23

In [LOS21, Example 5.10], the K\mathrm{K}-theory of the reduced group C\mathrm{C}^{*}-algebra of the group B4\mathrm{B}_{4} is computed to be

K0(Cred(B4))(/2),K1(Cred(B4)).\mathrm{K}_{0}(\mathrm{C}^{*}_{\mathrm{red}}(\mathrm{B}_{4}))\simeq\mathbb{Z}\oplus(\mathbb{Z}/2\mathbb{Z}),\qquad\qquad\mathrm{K}_{1}(\mathrm{C}^{*}_{\mathrm{red}}(\mathrm{B}_{4}))\simeq\mathbb{Z}.

The abelian group K1(Cred(B4))\mathrm{K}_{1}(\mathrm{C}^{*}_{\mathrm{red}}(\mathrm{B}_{4})) is generated by the element [λσ1]=[λσ2]=[λσ3][\lambda_{\sigma_{1}}]=[\lambda_{\sigma_{2}}]=[\lambda_{\sigma_{3}}]. It is not difficult to compute the index pairing.

9.6 Bi-orderings of the Thompson group F\mathrm{F}

The Thompson group F\mathrm{F} consists of orientation-preserving piecewise linear homeomorphisms ff of the interval [0,1][0,1] such that

  1. 1.

    there are only finitely many points of non-differentiability of ff,

  2. 2.

    each point of non-differentiability of ff occurs at a dyadic rational number a2n\frac{a}{2^{n}} where a,na,n\in\mathbb{N},

  3. 3.

    the derivatives on intervals of differentiability are powers of 2.

See Figure 1 for an example of element of this group. We refer to [CFP96] for an introduction to this group.

[Uncaptioned image]

Figure 1: an element of Thompson group F\mathrm{F}

A description of all bi-orderings on Thompson group is given in the paper [NaR10]. We describe four of them. Let fFf\in\mathrm{F} with fIdf\not=\mathrm{Id}. We consider the corresponding lateral derivatives ff^{\prime}_{-} and f+f^{\prime}_{+}. We denote by xfx^{-}_{f} the leftmost point xx for which f+(x)1f^{\prime}_{+}(x)\neq 1 and we denote by xf+x^{+}_{f} the rightmost point xx for which f(x)1f^{\prime}_{-}(x)\neq 1. In this context, we can define

  1. 1.

    the bi-ordering x+\preceq_{x^{-}}^{+} for which Idf\mathrm{Id}\prec f if and only if f+(xf)>1f^{\prime}_{+}(x^{-}_{f})>1,

  2. 2.

    the bi-ordering x\preceq_{x^{-}}^{-} for which Idf\mathrm{Id}\prec f if and only if f+(xf)<1f^{\prime}_{+}(x^{-}_{f})<1,

  3. 3.

    the bi-ordering x++\preceq_{x^{+}}^{+} for which Idf\mathrm{Id}\prec f if and only if f(xf+)<1f^{\prime}_{-}(x^{+}_{f})<1,

  4. 4.

    the bi-ordering x+\preceq_{x^{+}}^{-} for which Idf\mathrm{Id}\prec f if and only if f(xf+)>1f^{\prime}_{-}(x^{+}_{f})>1.

Our results pave the way for a better understanding of the K-theory and K-homology groups of Thompson group, whose properties remain largely unknown.

10 Banach Fredholm module associated to the free Hilbert transform

Let G=𝔽G=\mathbb{F}_{\infty} be the free group with a countable sequence of generators g1,g2,g_{1},g_{2},\ldots. For any integer n1n\geqslant 1, let Ln+:L2(VN(𝔽))L2(VN(𝔽))L_{n}^{+}\colon\mathrm{L}^{2}(\mathrm{VN}(\mathbb{F}_{\infty}))\to\mathrm{L}^{2}(\mathrm{VN}(\mathbb{F}_{\infty})) and Ln:L2(VN(𝔽))L2(VN(𝔽))L_{n}^{-}\colon\mathrm{L}^{2}(\mathrm{VN}(\mathbb{F}_{\infty}))\to\mathrm{L}^{2}(\mathrm{VN}(\mathbb{F}_{\infty})) be the orthogonal projections such that

Ln+(λs)={λss starts with the letter gn0otherwise,Ln(λs)={λss starts with the letter gn10otherwise.L_{n}^{+}(\lambda_{s})=\begin{cases}\lambda_{s}&s\text{ starts with the letter }g_{n}\\ 0&\text{otherwise}\end{cases},\quad L_{n}^{-}(\lambda_{s})=\begin{cases}\lambda_{s}&s\text{ starts with the letter }g_{n}^{-1}\\ 0&\text{otherwise}\end{cases}.

Let us further consider signs εn+,εn{1,1}\varepsilon_{n}^{+},\varepsilon_{n}^{-}\in\{-1,1\} for any integer n1n\geqslant 1. Following [MeR17], we define the free Hilbert transform associated with ε=(εn±)\varepsilon=(\varepsilon_{n}^{\pm}) as

(10.1) ε=defn1(εn+Ln++εnLn).\mathcal{H}_{\varepsilon}\overset{\mathrm{def}}{=}\sum_{n\geqslant 1}\big{(}\varepsilon_{n}^{+}L_{n}^{+}+\varepsilon_{n}^{-}L_{n}^{-}\big{)}.

Clearly, since the ranges of the various projections Ln±L_{n}^{\pm} are mutually orthogonal, the linear map ε\mathcal{H}_{\varepsilon} is bounded on the Hilbert space L2(VN(𝔽))\mathrm{L}^{2}(\mathrm{VN}(\mathbb{F}_{\infty})). The far-reaching generalization in [MeR17, Section 4] is that the map ε\mathcal{H}_{\varepsilon} induces a completely bounded map on the Banach space Lp(VN(𝔽))\mathrm{L}^{p}(\mathrm{VN}(\mathbb{F}_{\infty})) for any 1<p<1<p<\infty.

In the sequel, we consider the contractive homomorphism π:Cred(𝔽)B(Lp(VN(𝔽))\pi\colon\mathrm{C}^{*}_{\mathrm{red}}(\mathbb{F}_{\infty})\to\mathrm{B}(\mathrm{L}^{p}(\mathrm{VN}(\mathbb{F}_{\infty})), aMaa\mapsto M_{a}, where Ma:Lp(VN(𝔽))Lp(VN(𝔽))M_{a}\colon\mathrm{L}^{p}(\mathrm{VN}(\mathbb{F}_{\infty}))\to\mathrm{L}^{p}(\mathrm{VN}(\mathbb{F}_{\infty})), yayy\mapsto ay is the multiplication operator by aa.

Proposition 10.1

Suppose that 1<p<1<p<\infty. Consider a family ε=(εn±)n1\varepsilon=(\varepsilon_{n}^{\pm})_{n\geqslant 1} of signs. Then (Lp(VN(𝔽)),ε)(\mathrm{L}^{p}(\mathrm{VN}(\mathbb{F}_{\infty})),\mathcal{H}_{\varepsilon}) is a Banach Fredholm module over the reduced group C\mathrm{C}^{*}-algebra Cred(𝔽)\mathrm{C}^{*}_{\mathrm{red}}(\mathbb{F}_{\infty}).

Proof : Let s,t𝔽s,t\in\mathbb{F}_{\infty}. Assume that ss begins with gkηsg_{k}^{\eta_{s}} and that tsts begins with glμtsg_{l}^{\mu_{ts}}, where ηs,μts{1,1}\eta_{s},\mu_{ts}\in\{-1,1\} (identified with {,+}\{-,+\}). We have

[ε,Mλt](λs)=εMλt(λs)Mλtε(λs)\displaystyle\left[\mathcal{H}_{\varepsilon},M_{\lambda_{t}}\right](\lambda_{s})=\mathcal{H}_{\varepsilon}M_{\lambda_{t}}(\lambda_{s})-M_{\lambda_{t}}\mathcal{H}_{\varepsilon}(\lambda_{s})
=(10.1)(n1εn+Ln++εnLn)λtsMλt(n1εn+Ln++εnLn)(λs)\displaystyle\overset{\eqref{free-Hilbert}}{=}\bigg{(}\sum_{n\geqslant 1}\varepsilon_{n}^{+}L_{n}^{+}+\varepsilon_{n}^{-}L_{n}^{-}\bigg{)}\lambda_{ts}-M_{\lambda_{t}}\bigg{(}\sum_{n\geqslant 1}\varepsilon_{n}^{+}L_{n}^{+}+\varepsilon_{n}^{-}L_{n}^{-}\bigg{)}(\lambda_{s})
=εlμtsλtsMλt(εkηsλs)=εlμtsλtsεkηsλts.\displaystyle=\varepsilon_{l}^{\mu_{ts}}\lambda_{ts}-M_{\lambda_{t}}(\varepsilon_{k}^{\eta_{s}}\lambda_{s})=\varepsilon_{l}^{\mu_{ts}}\lambda_{ts}-\varepsilon_{k}^{\eta_{s}}\lambda_{ts}.

The other cases where s=es=e or ts=ets=e are left to the reader. We deduce that the rank of the commutator [ε,Mλt]\left[\mathcal{H}_{\varepsilon},M_{\lambda_{t}}\right] is zero or one. If a=s𝔽asλsa=\sum_{s\in\mathbb{F}_{\infty}}a_{s}\lambda_{s} is a finite sum, where each asa_{s} is a complex number, belonging to the group algebra [𝔽]\mathbb{C}[\mathbb{F}_{\infty}] then we see by linearity that the commutator [,Ma]\left[\mathcal{H},M_{a}\right] is a finite-rank operator. We infer that for any aCred(𝔽)a\in\mathrm{C}^{*}_{\mathrm{red}}(\mathbb{F}_{\infty}) the commutator [,Ma]\left[\mathcal{H},M_{a}\right] is compact, as a limit of finite-rank operators, since the homomorphism π:Cred(𝔽)B(Lp(VN(𝔽)))\pi\colon\mathrm{C}^{*}_{\mathrm{red}}(\mathbb{F}_{\infty})\to\mathrm{B}(\mathrm{L}^{p}(\mathrm{VN}(\mathbb{F}_{\infty}))), aMaa\mapsto M_{a} is continuous.

Note that ε(λe)=0\mathcal{H}_{\varepsilon}(\lambda_{e})=0. Consequently, with (10.1) it is easy to check that ε2IdLp(VN(𝔽)\mathcal{H}_{\varepsilon}^{2}\sim\mathrm{Id}_{\mathrm{L}^{p}(\mathrm{VN}(\mathbb{F}_{\infty})}.    

Remark 10.2

By restriction, we can obtain a Banach Fredholm module over each reduced group C\mathrm{C}^{*}-algebra Cred(𝔽n)\mathrm{C}^{*}_{\mathrm{red}}(\mathbb{F}_{n}) for any integer n1n\geqslant 1. By [PiV82, Corollary 3.2 p. 152], we have K0(Cred(𝔽n))=\mathrm{K}_{0}(\mathrm{C}^{*}_{\mathrm{red}}(\mathbb{F}_{n}))=\mathbb{Z} and the generator of K0(Cred(𝔽n))\mathrm{K}_{0}(\mathrm{C}^{*}_{\mathrm{red}}(\mathbb{F}_{n})) is [1][1]. The same reference gives an isomorphism K1(Cred(𝔽n))=n\mathrm{K}_{1}(\mathrm{C}^{*}_{\mathrm{red}}(\mathbb{F}_{n}))=\mathbb{Z}^{n} and the generators of the group K1(Cλ(𝔽n))\mathrm{K}_{1}(\mathrm{C}^{*}_{\lambda}(\mathbb{F}_{n})) are [λg1],,[λgn][\lambda_{g_{1}}],\ldots,[\lambda_{g_{n}}]. We invite the reader to determine the index pairing.

11 Semigroups, Fredholm modules and vectorial Riesz transforms

Hilbert bimodules

We start by reviewing the concept of Hilbert bimodule which is crucial for defining the derivations that allow the introduction of the Hodge-Dirac operators considered in this paper. Let \mathcal{M} be a von Neumann algebra. A Hilbert \mathcal{M}-bimodule is a Hilbert space \mathcal{H} together with a *-representation Φ:B()\Phi\colon\mathcal{M}\to\mathrm{B}(\mathcal{H}) and a *-anti-representation Ψ:B()\Psi\colon\mathcal{M}\to\mathrm{B}(\mathcal{H}) such that Φ(x)Ψ(y)=Ψ(y)Φ(x)\Phi(x)\Psi(y)=\Psi(y)\Phi(x) for any x,yx,y\in\mathcal{M}. For all x,yx,y\in\mathcal{M} and any ξ\xi\in\mathcal{H}, we let xξy=defΦ(x)Ψ(y)ξx\xi y\overset{\mathrm{def}}{=}\Phi(x)\Psi(y)\xi. We say that the bimodule is normal if Φ\Phi and Ψ\Psi are normal, i.e. weak* continuous. The bimodule is said to be symmetric if there exists an antilinear involution 𝒥:\mathcal{J}\colon\mathcal{H}\to\mathcal{H} such that 𝒥(xξy)=y𝒥(ξ)x\mathcal{J}(x\xi y)=y^{*}\mathcal{J}(\xi)x^{*} for any x,yx,y\in\mathcal{M} and any ξ\xi\in\mathcal{H}.

W\mathrm{W}^{*}-derivations

Now, we introduce a notion of derivation that can be viewed as an abstract version of a gradient. If (,Φ,Ψ)(\mathcal{H},\Phi,\Psi) is a Hilbert \mathcal{M}-bimodule, then following [Wea96, p. 267] we define a W\mathrm{W}^{*}-derivation to be a weak* closed densely defined unbounded operator :dom\partial\colon\operatorname{dom}\partial\subset\mathcal{M}\to\mathcal{H} such that the domain dom\operatorname{dom}\partial is a weak* dense unital *-subalgebra of \mathcal{M} and

(11.1) (fg)=f(g)+(f)g,f,gdom.\partial(fg)=f\partial(g)+\partial(f)g,\quad f,g\in\operatorname{dom}\partial.

We say that a W\mathrm{W}^{*}-derivation is symmetric if the bimodule (,Φ,Ψ)(\mathcal{H},\Phi,\Psi) is symmetric and if we have 𝒥((f))=(f)\mathcal{J}(\partial(f))=\partial(f^{*}) for any xdomx\in\operatorname{dom}\partial. In the sequel, we let =defdom\mathcal{B}\overset{\mathrm{def}}{=}\operatorname{dom}\partial.

The triple (L(),L2()2,D)(\mathrm{L}^{\infty}(\mathcal{M}),\mathrm{L}^{2}(\mathcal{M})\oplus_{2}\mathcal{H},D)

From a derivation, we will now explain how to introduce a triple in the spirit of noncommutative geometry. Let :dom\partial\colon\operatorname{dom}\partial\subset\mathcal{M}\to\mathcal{H} be a W\mathrm{W}^{*}-derivation where the von Neumann algebra \mathcal{M} is equipped with a normal faithful finite trace τ\tau. Suppose that the operator :domL2()\partial\colon\operatorname{dom}\partial\subset\mathrm{L}^{2}(\mathcal{M})\to\mathcal{H} is closable. We denote again its closure by \partial. Note that the subspace \mathcal{B} is a core of \partial. Recall that it is folklore and well-known that a weak* dense subalgebra of \mathcal{M} is dense in the space L2()\mathrm{L}^{2}(\mathcal{M}). As the operator \partial is densely defined and closed, by [Kat76, Theorem 5.29 p. 168] the adjoint operator \partial^{*} is densely defined and closed on L2()\mathrm{L}^{2}(\mathcal{M}) and =\partial^{**}=\partial.

Following essentially [HiT13b] and [Cip16], we introduce the unbounded closed operator DD on the Hilbert space L2()2\mathrm{L}^{2}(\mathcal{M})\oplus_{2}\mathcal{H} introduced in (11.6) and defined by

(11.2) D(f,g)=def((g),(f)),fdom,gdom.D(f,g)\overset{\mathrm{def}}{=}\big{(}\partial^{*}(g),\partial(f)\big{)},\quad f\in\operatorname{dom}\partial,\ g\in\operatorname{dom}\partial^{*}.

We call it the Hodge-Dirac operator associated to \partial and it can be written as in (11.6). It is not difficult to check that this operator is selfadjoint. If fL()f\in\mathrm{L}^{\infty}(\mathcal{M}), we define the bounded operator πf:L2()2L2()2\pi_{f}\colon\mathrm{L}^{2}(\mathcal{M})\oplus_{2}\mathcal{H}\to\mathrm{L}^{2}(\mathcal{M})\oplus_{2}\mathcal{H} by

(11.3) πf=def[Mf00Φf],fL()\pi_{f}\overset{\mathrm{def}}{=}\begin{bmatrix}\mathrm{M}_{f}&0\\ 0&\Phi_{f}\\ \end{bmatrix},\quad f\in\mathrm{L}^{\infty}(\mathcal{M})

where the linear map Mf:L2()L2()\mathrm{M}_{f}\colon\mathrm{L}^{2}(\mathcal{M})\to\mathrm{L}^{2}(\mathcal{M}), gfgg\mapsto fg is the multiplication operator by ff and where Φf:\Phi_{f}\colon\mathcal{H}\to\mathcal{H}, hfhh\mapsto fh is the left bimodule action.

Note that completely Dirichlet forms give rise to W\mathrm{W}^{*}-derivations, see [CiS03], [Cip97], [Cip08], [Cip16] and [Wir22] and references therein. More precisely, if \mathcal{E} is a completely Dirichlet form on a noncommutative L2\mathrm{L}^{2}-space L2()\mathrm{L}^{2}(\mathcal{M}) with associated semigroup (Tt)t0(T_{t})_{t\geqslant 0} and associated operator A2A_{2}, there exist a symmetric Hilbert \mathcal{M}-bimodule (,Φ,Ψ,𝒥)(\mathcal{H},\Phi,\Psi,\mathcal{J}) and a W\mathrm{W}^{*}-derivation :dom\partial\colon\operatorname{dom}\partial\subset\mathcal{M}\to\mathcal{H} such that :domL2()\partial\colon\operatorname{dom}\partial\subset\mathrm{L}^{2}(\mathcal{M})\to\mathcal{H} is closable with closure also denoted by \partial and satisfying A2=A_{2}=\partial^{*}\partial. In other words, this means that the opposite of the infinitesimal generator A2-A_{2} of a symmetric sub-Markovian semigroup of operators can always be represented as the composition of a ¡¡divergence¿¿ \partial^{*} with a ¡¡gradient¿¿ \partial. Note that the symmetric Hilbert \mathcal{M}-bimodule (,Φ,Ψ,𝒥)(\mathcal{H},\Phi,\Psi,\mathcal{J}) is not unique since one can always artificially expand \mathcal{H}. However, if we add an additional condition, namely that ¡¡the bimodule is generated by the derivation \partial¿¿, we obtain uniqueness, see [Wir22, Theorem 6.9] for a precise statement. In this case, the bimodule (and its associated derivation) is called the tangent bimodule associated with the completely Dirichlet form (or the associated semigroup).

This fundamental result allows anyone to introduce a triple (L(Ω),L2(Ω)2,D)(\mathrm{L}^{\infty}(\Omega),\mathrm{L}^{2}(\Omega)\oplus_{2}\mathcal{H},D) associated to the semigroup in the spirit of noncommutative geometry. Here DD is the unbounded selfadjoint operator acting on a dense subspace of the Hilbert space L2(Ω)2\mathrm{L}^{2}(\Omega)\oplus_{2}\mathcal{H} defined by

(11.4) D=def[00].D\overset{\mathrm{def}}{=}\begin{bmatrix}0&\partial^{*}\\ \partial&0\end{bmatrix}.

The Hodge-Dirac operator DD of (11.6) is related to the operator A2A_{2} by

(11.5) D2=(11.6)[00]2=[00]=[A200].D^{2}\overset{\eqref{Hodge-Dirac-I}}{=}\begin{bmatrix}0&\partial^{*}\\ \partial&0\end{bmatrix}^{2}=\begin{bmatrix}\partial^{*}\partial&0\\ 0&\partial\partial^{*}\end{bmatrix}=\begin{bmatrix}A_{2}&0\\ 0&\partial\partial^{*}\end{bmatrix}.

Now, suppose that 1<p<1<p<\infty. Sometimes, the map 2\partial_{2} induces a closable unbounded operator :domLp()𝒳p\partial\colon\operatorname{dom}\partial\subset\mathrm{L}^{p}(\mathcal{M})\to\mathcal{X}_{p} for some Banach space 𝒳p\mathcal{X}_{p}. Denoting by p\partial_{p} its closure, we can consider the Lp\mathrm{L}^{p}-realization of the previous operator

(11.6) Dp=def[0(p)p0]D_{p}\overset{\mathrm{def}}{=}\begin{bmatrix}0&(\partial_{p^{*}})^{*}\\ \partial_{p}&0\end{bmatrix}

as acting on a dense subspace of the Banach space Lp()p𝒳p\mathrm{L}^{p}(\mathcal{M})\oplus_{p}\mathcal{X}_{p}. If the operator DpD_{p} is bisectorial and admits a bounded H(Σθbi)\mathrm{H}^{\infty}(\Sigma_{\theta}^{\mathrm{bi}}) functional calculus for some θ(0,π2)\theta\in(0,\frac{\pi}{2}), we have

signDp=Dp|Dp|1=Dp(Dp2)12=[0(p)p0]([Ap00p(p)])12=[0pAp120].\displaystyle\operatorname{\mathrm{sign}}D_{p}=D_{p}|D_{p}|^{-1}=D_{p}(D_{p}^{2})^{-\frac{1}{2}}=\begin{bmatrix}0&(\partial_{p^{*}})^{*}\\ \partial_{p}&0\end{bmatrix}\left(\begin{bmatrix}A_{p}&0\\ 0&\partial_{p}(\partial_{p^{*}})^{*}\end{bmatrix}\right)^{-\frac{1}{2}}=\begin{bmatrix}0&*\\ \partial_{p}A_{p}^{-\frac{1}{2}}&0\end{bmatrix}.

In particular, the operator pAp12:Lp()𝒳p\partial_{p}A_{p}^{-\frac{1}{2}}\colon\mathrm{L}^{p}(\mathcal{M})\to\mathcal{X}_{p} is bounded.

Definition 11.1

We say that =defpAp12\mathcal{R}\overset{\mathrm{def}}{=}\partial_{p}A_{p}^{-\frac{1}{2}} is the (abstract) vectorial Riesz transform associated with the symmetric sub-Markovian semigroup (Tt)t0(T_{t})_{t\geqslant 0}.

From the perspective of the action of the algebra or the kernel of the Dirac operator, the Banach space 𝒳p\mathcal{X}_{p} is sometimes too large and needs to be replaced by a smaller space. In this case, we consider the closed subspace

(11.7) Ωp=defspan{gp(f):fdomp,gL()¯}\Omega_{p}\overset{\mathrm{def}}{=}\overline{\mathrm{span}\big{\{}g\partial_{p}(f):f\in\operatorname{dom}\partial_{p},g\in\mathrm{L}^{\infty}(\mathcal{M})}\big{\}}

of the Banach space 𝒳p\mathcal{X}_{p}. And we replace the Banach space 𝒳p\mathcal{X}_{p} by the subspace Ωp\Omega_{p}. The advantage of this space is the following result.

Proposition 11.2

The Banach space Ωp\Omega_{p} is a L()\mathrm{L}^{\infty}(\mathcal{M})-bimodule.

Proof : For any fdomf\in\operatorname{dom}\partial and any g,hL()g,h\in\mathrm{L}^{\infty}(\mathcal{M}), note that

g(h(f))=gh(f).g(h\partial(f))=gh\partial(f).

Thus by linearity and density, Ωp\Omega_{p} is a right L()\mathrm{L}^{\infty}(\mathcal{M})-module. Moreover, for any gL()g\in\mathrm{L}^{\infty}(\mathcal{M}) and any f,hdomf,h\in\operatorname{dom}\partial, we have

g(f)h=(11.1)g[(fh)f(h)]=g(hf)gf(h).\displaystyle g\partial(f)h\overset{\eqref{Leibniz}}{=}g\big{[}\partial(fh)-f\partial(h)\big{]}=g\partial(hf)-gf\partial(h).

Thus g(f)hg\partial(f)h belongs to Ωp\Omega_{p}. Since dom\operatorname{dom}\partial is a core for p\partial_{p}, the same holds for fdompf\in\operatorname{dom}\partial_{p}. If hL()h\in\mathrm{L}^{\infty}(\mathcal{M}) is a general element, we approximate it in the strong operator topology by a bounded net in dom\operatorname{dom}\partial and obtain again that the same holds for hL()h\in\mathrm{L}^{\infty}(\mathcal{M}). By linearity and density, we deduce that Ωp\Omega_{p} is a left L()\mathrm{L}^{\infty}(\mathcal{M})-module, so finally a L()\mathrm{L}^{\infty}(\mathcal{M})-bimodule.    

Remark 11.3

It is interesting to observe that if L()\mathrm{L}^{\infty}(\mathcal{M}) is commutative, i.e. the algebra L(Ω)\mathrm{L}^{\infty}(\Omega) of a (localizable) measure space Ω\Omega, the previous bimodule is not necessarily commutative, i.e. we does not have gkh=hkggkh=hkg for any g,hL(Ω)g,h\in\mathrm{L}^{\infty}(\Omega) and kΩpk\in\Omega_{p}. In our previous paper, we referred to this phenomenon as ¡¡hidden noncommutative geometry¿¿.

If fL()f\in\mathrm{L}^{\infty}(\mathcal{M}), we define the bounded operator πf:Lp()pΩpLp()pΩp\pi_{f}\colon\mathrm{L}^{p}(\mathcal{M})\oplus_{p}\Omega_{p}\to\mathrm{L}^{p}(\mathcal{M})\oplus_{p}\Omega_{p} by

(11.8) πf=def[Mf00Φf],fL()\pi_{f}\overset{\mathrm{def}}{=}\begin{bmatrix}\mathrm{M}_{f}&0\\ 0&\Phi_{f}\\ \end{bmatrix},\quad f\in\mathrm{L}^{\infty}(\mathcal{M})

where the linear map Mf:Lp()Lp()\mathrm{M}_{f}\colon\mathrm{L}^{p}(\mathcal{M})\to\mathrm{L}^{p}(\mathcal{M}), gfgg\mapsto fg is the multiplication operator by ff and where Φf:ΩpΩp\Phi_{f}\colon\Omega_{p}\to\Omega_{p}, hfhh\mapsto fh is the left bimodule action provided by Proposition 11.2.

If (Lp()pΩp,signDp,[Id00Id])\bigg{(}\mathrm{L}^{p}(\mathcal{M})\oplus_{p}\Omega_{p},\operatorname{\mathrm{sign}}D_{p},\begin{bmatrix}-\mathrm{Id}&0\\ 0&\mathrm{Id}\\ \end{bmatrix}\bigg{)} is a (well-defined) even Banach Fredholm module over a subalgebra 𝒜\mathcal{A} of L()\mathrm{L}^{\infty}(\mathcal{M}), the pairing with the K-homology group K0(𝒜)\mathrm{K}_{0}(\mathcal{A}) is given by the following formula. If eMn(𝒜)e\in\mathrm{M}_{n}(\mathcal{A}) is an idempotent then

(11.9) [e],(Lp()pΩp,signD)K0(𝒜),K0(𝒜,ncp)=(4.6)Indexen(Id)en.\big{\langle}[e],(\mathrm{L}^{p}(\mathcal{M})\oplus_{p}\Omega_{p},\operatorname{\mathrm{sign}}D)\big{\rangle}_{\mathrm{K}_{0}(\mathcal{A}),\mathrm{K}^{0}(\mathcal{A},\mathscr{L}^{p}_{\mathrm{nc}})}\overset{\eqref{pairing-even-1}}{=}\operatorname{Index}e_{n}(\mathrm{Id}\otimes\mathcal{R})e_{n}.
Remark 11.4

Note that with Proposition 3.13, we can obtain sometimes an odd Banach Fredholm module with signDp\operatorname{\mathrm{sign}}D_{p}, so a pairing with the group K1(𝒜)\mathrm{K}_{1}(\mathcal{A}).

12 Future directions

We plan to elaborate on the ideas presented in the final section in a future update of this very preliminary preprint. We also plan to expand some sections and explain the analogies with [ScS22]. Finally, we will conduct an exhaustive study of the groups K0(𝒜,)\mathrm{K}^{0}(\mathcal{A},\mathscr{B}) and K1(𝒜,)\mathrm{K}^{1}(\mathcal{A},\mathscr{B}) (and some generalizations) in a forthcoming paper.

Declaration of interest

None.

Competing interests

The author declares that he has no competing interests.

Data availability

No data sets were generated during this study.

Acknowledgment

The author would like to thank Hermann Schulz-Baldes for bringing his recent book [ScS22] to our attention on his own initiative.

References

  • [AbA02] Y. A. Abramovich and C. D. Aliprantis. An invitation to operator theory. Graduate Studies in Mathematics, 50. American Mathematical Society, Providence, RI, 2002.
  • [ADM96] D. Albrecht, X. Duong and A. McIntosh. Operator theory and harmonic analysis. Instructional Workshop on Analysis and Geometry, Part III (Canberra, 1995), 77–136. Proc. Centre Math. Appl. Austral. Nat. Univ., 34. Australian National University, Centre for Mathematics and its Applications, Canberra, 1996.
  • [AmG16] B. Ammann and N. Große. LpL^{p}-spectrum of the Dirac operator on products with hyperbolic spaces. Calc. Var. Partial Differential Equations 55 (2016), no. 5, Art. 127.
  • [ArK22] C. Arhancet and C. Kriegler. Riesz transforms, Hodge-Dirac operators and functional calculus for multipliers. Lecture Notes in Mathematics, 2304. Springer, Cham, 2022.
  • [Arh24a] C. Arhancet. Sobolev algebras on Lie groups and noncommutative geometry. J. Noncommut. Geom. 18 (2024), no. 2, 451–500.
  • [Arh24b] C. Arhancet. Spectral triples, Coulhon-Varopoulos dimension and heat kernel estimates. Adv. Math. 451 (2024), Paper No. 109794.
  • [Arh24c] C. Arhancet. Curvature, Hodge-Dirac operators and Riesz transforms. Preprint, arXiv:2401.04203.
  • [Arh24d] C. Arhancet. Riesz transforms and approximation numbers on noncommutative tori. Preprint.
  • [ABG90] N. Asmar, E. Berkson and T. A. Gillespie. Representations of groups with ordered duals and generalized analyticity. J. Funct. Anal. 90 (1990), no. 1, 206–235.
  • [Arv67] W. B. Arveson. Analyticity in operator algebras. Amer. J. Math. 89 (1967), 578–642.
  • [AIM09] K. Astala, T. Iwaniec and G. Martin. Elliptic partial differential equations and quasiconformal mappings in the plane. Princeton Math. Ser., 48. Princeton University Press, Princeton, NJ, 2009.
  • [AMN97] P. Auscher, A. McIntosh, A. Nahmod. The square root problem of Kato in one dimension, and first order elliptic systems. Indiana Univ. Math. J. 46 (1997), no. 3, 659–695.
  • [AHLMT02] P. Auscher, S. Hofmann, M. Lacey, A. McIntosh and P. Tchamitchian. The solution of the Kato square root problem for second order elliptic operators on n\mathbb{R}^{n}. Ann. of Math. (2) 156 (2002), no. 2, 633–654.
  • [AKM06] A. Axelsson, S. Keith and A. McIntosh. Quadratic estimates and functional calculi of perturbed Dirac operators. Invent. Math. 163 (2006), no. 3, 455–497.
  • [ABARW22] S. Azzali, Sarah L. Browne, Maria Paula Gomez Aparicio, Lauren C. Ruth, Hang Wang. K-homology and K-theory of pure Braid groups. New York J. Math. 28 (2022), 1256–1294.
  • [BBV99] C. Béguin, H. Bettaieb and A. Valette. K-theory for CC^{*}-algebras of one-relator groups. K-Theory 16 (1999), no. 3, 277–298.
  • [Bl98] B. Blackadar. KK-theory for operator algebras. Math. Sci. Res. Inst. Publ., 5. Cambridge University Press, Cambridge, 1998.
  • [Bar07] C. Bär. Conformal structures in noncommutative geometry. J. Noncommut. Geom. 1 (2007), no. 3, 385–395.
  • [Bek15] T. N. Bekjan. Noncommutative Hardy space associated with semi-finite subdiagonal algebras. J. Math. Anal. Appl. 429 (2015), no. 2, 1347–1369.
  • [BEO13] A. Bényi and T. Oh. The Sobolev inequality on the torus revisited. Publ. Math. Debrecen 83 (2013), no. 3, 359–374.
  • [BlL07] D. Blecher and L. E. Labuschagne. Von Neumann algebraic HpH^{p} theory. Function spaces, 89–114. Contemp. Math., 435. American Mathematical Society, Providence, RI, 2007.
  • [BlB13] D. D. Bleecker and B. Booß-Bavnbek. Index theory—with applications to mathematics and physics. International Press, Somerville, MA, 2013.
  • [BoS06] A. Böttcher and B. Silbermann. Analysis of Toeplitz operators. Springer-Verlag, Berlin, 2006.
  • [Cal77] A.-P. Calderón. Cauchy integrals on Lipschitz curves and related operators. Proc. Nat. Acad. Sci. U.S.A. 74 (1977), no. 4, 1324–1327.
  • [Cam17] M. C. Câmara. Toeplitz operators and Wiener-Hopf factorisation: an introduction. Concr. Oper. 4 (2017), no. 1, 130–145.
  • [CFP96] J. Cannon and W. Floyd and W. Parry. Introductory notes on Richard Thompson’s groups. L’Enseignement Mathématique 42 (1996), 215–256.
  • [CPR11] A. L. Carey, J. Phillips and A. Rennie. Spectral triples: examples and index theory. ESI Lect. Math. Phys. European Mathematical Society (EMS), Zürich, 2011, 175–265.
  • [CGRS14] A. L. Carey, V. Gayral, A. Rennie and F. A. Sukochev. Index theory for locally compact noncommutative geometries. Mem. Amer. Math. Soc. 231 (2014), no. 1085.
  • [Cas22] R. E. Castillo. Fourier meets Hilbert and Riesz—an introduction to the corresponding transforms. De Gruyter Stud. Math., 87. De Gruyter, Berlin, 2022.
  • [CXY13] Z. Chen, Q. Xu and Z. Yin. Harmonic analysis on quantum tori. Comm. Math. Phys. 322 (2013), no. 3, 755–805.
  • [Cip08] F. Cipriani. Dirichlet forms on noncommutative spaces. Quantum potential theory, 161–276, Lecture Notes in Math., 1954, Springer, Berlin, 2008.
  • [Cip16] F. Cipriani. Noncommutative potential theory: a survey. J. Geom. Phys. 105 (2016), 25–59.
  • [CiS03] F. Cipriani and J.-L. Sauvageot. Derivations as square roots of Dirichlet forms. J. Funct. Anal. 201 (2003), no. 1, 78–120.
  • [CGIS14] F. Cipriani, D. Guido, T. Isola and J.-L. Sauvageot. Spectral triples for the Sierpinski gasket. J. Funct. Anal. 266 (2014), no. 8, 4809–4869.
  • [Cip97] F. Cipriani. Dirichlet forms and Markovian semigroups on standard forms of von Neumann algebras. J. Funct. Anal. 147 (1997), no. 2, 259–300.
  • [ClR16] A. Clay and D. Rolfsen. Ordered groups and topology. Grad. Stud. Math., 176. American Mathematical Society, Providence, RI, 2016.
  • [CMM82] R. R. Coifman, A. McIntosh and Y. Meyer. L’intégrale de Cauchy définit un opérateur borné sur L2L^{2} pour les courbes lipschitziennes. Ann. of Math. (2) 116 (1982), no. 2, 361–387.
  • [Con94] A. Connes. Noncommutative geometry. Academic Press, Inc., San Diego, CA, 1994.
  • [CoM95] A. Connes and H. Moscovici. The local index formula in noncommutative geometry. Geom. Funct. Anal. 5 (1995), no. 2, 174–243.
  • [CoM08] A. Connes and M. Marcolli. A walk in the noncommutative garden. An invitation to noncommutative geometry, 1–128, World Sci. Publ., Hackensack, NJ, 2008.
  • [Dav89] E. B. Davies. Heat kernels and spectral theory. Cambridge Tracts in Mathematics, 92. Cambridge University Press, Cambridge, 1989.
  • [Deh94] P. Dehornoy. Braid groups and left distributive operations. Trans. Amer. Math. Soc. 345 (1994), no. 1, 115–150.
  • [DDRW08] P. Dehornoy, I. Dynnikov, D. Rolfsen and B. Wiest. Ordering braids. Surveys and Monographs, vol. 148, American Mathematical Society, Providence, RI, 2008.
  • [DNR16] B. Deroin, A. Navas, C. Rivas. Groups, Orders, and Dynamics. Preprint, arXiv:1408.5805.
  • [EcI18] M. Eckstein and B. Iochum. Spectral action in noncommutative geometry. SpringerBriefs in Mathematical Physics, 27. Springer, Cham, 2018.
  • [Ege15] M. Egert. On Kato’s conjecture and mixed boundary conditions. PhD, 2015.
  • [EnN00] K.-J. Engel and R. Nagel. One-parameter semigroups for linear evolution equations. Graduate Texts in Mathematics, 194. Springer-Verlag, New York, 2000.
  • [EmN18] H. Emerson and B. Nica. K-homological finiteness and hyperbolic group. J. Reine Angew. Math. 745 (2018), 189–229.
  • [FGMR19] I. Forsyth, M. Goffeng, B. Mesland and A. Rennie. Boundaries, spectral triples and K-homology. J. Noncommut. Geom. 13 (2019), no. 2, 407–472.
  • [Ger22] D. M. Gerontogiannis. On finitely summable Fredholm modules from Smale spaces. Trans. Amer. Math. Soc. 375 (2022), no. 12, 8885–8944.
  • [GPX22] A. Gonzalez-Perez, J. Parcet, R. Xia. Noncommutative Cotlar identities for groups acting on tree-like structures. Preprint, arXiv:2209.05298.
  • [GVF01] J. M. Gracia-Bondia, J. C. Varilly and H. Figueroa. Elements of noncommutative geometry. Birkhäuser Advanced Texts: Basler Lehrbücher. Birkhäuser Boston, Inc., Boston, MA, 2001.
  • [Gra14b] L. Grafakos. Modern Fourier analysis. Third edition. Grad. Texts in Math., 250. Springer, New York, 2014.
  • [Haa05] M. Haase. Spectral mapping theorems for holomorphic functional calculi. J. London Math. Soc. (2) 71 (2005), no.3, 723–739.
  • [Haa06] M. Haase. The functional calculus for sectorial operators. Operator Theory: Advances and Applications, 169. Birkhäuser Verlag (2006).
  • [Had03] T. Hadfield. The noncommutative geometry of the discrete Heisenberg group. Houston J. Math. 29 (2003), no. 2, 453–481.
  • [Had04] T. Hadfield. Fredholm modules over certain group CC^{*}-algebras. J. Operator Theory 51 (2004), no. 1, 141–160.
  • [HiR00] N. Higson and J. Roe. Analytic K-homology. Oxford Math. Monogr. Oxford Sci. Publ. Oxford University Press, Oxford, 2000.
  • [Hig02] N. Higson. The Local Index Formula in Noncommutative Geometry. Lectures given at the School and Conference on Algebraic K-theory and Its Applications. Trieste, 8-26 July 2002.
  • [HiT13b] M. Hinz and T. Teplyaev. Vector analysis on fractals and applications. Fractal geometry and dynamical systems in pure and applied mathematics. II. Fractals in applied mathematics, 147–163, Contemp. Math., 601, Amer. Math. Soc., Providence, RI, 2013.
  • [HLM02] S. Hofmann, M. Lacey and A. McIntosh. The solution of the Kato problem for divergence form elliptic operators with Gaussian heat kernel bounds. Ann. of Math. (2) 156 (2002), no. 2, 623–631.
  • [HvNVW16] T. Hytönen, J. van Neerven, M. Veraar and L. Weis. Analysis in Banach spaces, Volume I: Martingales and Littlewood-Paley theory. Springer, 2016.
  • [HvNVW18] T. Hytönen, J. van Neerven, M. Veraar and L. Weis. Analysis in Banach spaces, Volume II: Probabilistic Methods and Operator Theory. Springer, 2018.
  • [HvNVW23] T. Hytönen, J. van Neerven, M. Veraar and L. Weis. Analysis in Banach spaces, Volume III: Harmonic Analysis and Spectral Theory. Springer, 2023.
  • [JaW82] S. Janson and T. H. Wolff. Schatten classes and commutators of singular integral operators. Ark. Mat. 20 (1982), no. 2, 301–310.
  • [Ji14] G. Ji. Analytic Toeplitz algebras and the Hilbert transform associated with a subdiagonal algebra. Sci. China Math. 57 (2014), no.3, 579–588.
  • [Jou83] J.-L. Journé. Calderón-Zygmund operators, pseudodifferential operators and the Cauchy integral of Calderón. Lecture Notes in Math., 994. Springer-Verlag, Berlin, 1983.
  • [KaT08] C. Kassel and V. Turaev. Braid groups. With the graphical assistance of Olivier Dodane. Grad. Texts in Math., 247. Springer, New York, 2008.
  • [Kat76] T. Kato. Perturbation theory for linear operators. Second edition. Grundlehren der Mathematischen Wissenschaften, Band 132. Springer-Verlag, Berlin-New York, 1976.
  • [Kha13] M. Khalkhali. Basic noncommutative geometry. EMS Ser. Lect. Math. European Mathematical Society (EMS), Zürich, 2013.
  • [Kin09a] F. King. Hilbert transforms. Vol. 1. Encyclopedia of Mathematics and its Applications, 124. Cambridge University Press, Cambridge, 2009.
  • [Kon80] H. König. ss-numbers, eigenvalues and the trace theorem in Banach spaces. Studia Math. 67 (1980), no. 2, 157–172.
  • [Kon86] H. König. Eigenvalue distribution of compact operators. Oper. Theory Adv. Appl., 16. Birkhäuser Verlag, Basel, 1986.
  • [Kon01] H. König. Eigenvalues of operators and applications. North-Holland Publishing Co., Amsterdam, 2001, 941–974.
  • [LaX13] L. E. Labuschagne and Q. Xu. A Helson-Szegö theorem for subdiagonal subalgebras with applications to Toeplitz operators. J. Funct. Anal. 265 (2013), no. 4, 545–561.
  • [LOS21] X. Li, T. Omland and J. Spielberg. CC^{*}-algebras of right LCM one-relator monoids and Artin-Tits monoids of finite type. Comm. Math. Phys. 381 (2021), no. 3, 1263–1308.
  • [LNWW20] J. Li, T. T. T. Nguyen, L. A. Ward and B. D. Wick. The Cauchy integral, bounded and compact commutators. Studia Math. 250 (2020), no. 2, 193–216.
  • [McQ91] A. McIntosh and T. Qian. Convolution singular integral operators on Lipschitz curves. Lecture Notes in Math. 1494. Springer-Verlag, Berlin, 1991, 142–162.
  • [MSX19] E. McDonald, F. Sukochev and X. Xiong. Quantum differentiability on quantum tori. Comm. Math. Phys. 371 (2019), no. 3, 1231–1260.
  • [MaW98] M. Marsalli and G. West. Noncommutative HpH^{p} spaces. J. Operator Theory 40 (1998), no. 2, 339–355.
  • [Mas09] J. Mashreghi. Representation theorems in Hardy spaces. London Math. Soc. Stud. Texts, 74. Cambridge University Press, Cambridge, 2009.
  • [MeR17] T. Mei and É. Ricard. Free Hilbert Transforms. Duke Math. J. 166 (2017), no. 11, 2153–2182.
  • [NaR10] A. Navas and C. Rivas. Describing all bi-orderings on Thompson’s group FF. Groups Geom. Dyn. 4 (2010), no. 1, 163–177.
  • [NeT11] S. Neshveyev and L. Tuset. KK-homology class of the Dirac operator on a compact quantum group. Doc. Math. 16 (2011), 767–780.
  • [Nik02] N. K. Nikolski. Operators, functions, and systems: an easy reading. Vol. 1. Math. Surveys Monogr., 92. American Mathematical Society, Providence, RI, 2002.
  • [Ouh05] E. M. Ouhabaz. Analysis of heat equations on domains. London Mathematical Society Monographs Series, 31. Princeton University Press, Princeton, NJ, 2005.
  • [Pie74] A. Pietsch. ss-numbers of operators in Banach spaces. Studia Math. 51 (1974), 201–223.
  • [Pi80] A. Pietsch. Operator ideals. North-Holland Math. Library, 20. North-Holland Publishing Co., Amsterdam-New York, 1980.
  • [Pie87] A. Pietsch. Eigenvalues and ss-numbers. Math. Anwendungen Phys. Tech., 43. Akademische Verlagsgesellschaft Geest & Portig K.-G., Leipzig, 1987.
  • [Pie07] A. Pietsch. History of Banach spaces and linear operators. Birkhäuser Boston, Inc., Boston, MA, 2007.
  • [Pie23] A. Pietsch. Traces on operator ideals defined over the class of all Banach spaces and related open problems. Integral Equations Operator Theory 95 (2023), no. 2, Paper No. 11.
  • [PiV82] M. Pimsner and D. Voiculescu. K-groups of reduced crossed products by free groups. J. Operator Theory 8 (1982), no. 1, 131–156.
  • [PiX03] G. Pisier and Q. Xu. Non-commutative LpL^{p}-spaces. In Handbook of the Geometry of Banach Spaces, Vol. II, edited by W.B. Johnson and J. Lindenstrauss, Elsevier, 1459–1517, 2003.
  • [PoV18] S. Pooya and A. Valette. K-theory for the CC^{*}-algebras of the solvable Baumslag-Solitar groups. Glasg. Math. J. 60 (2018), no.2, 481–486.
  • [PSB16] E. Prodan and H. Schulz-Baldes. Bulk and boundary invariants for complex topological insulators. Math. Phys. Stud. Springer, [Cham], 2016.
  • [Pus11] M. Puschnigg. Finitely summable Fredholm modules over higher rank groups and lattices. J. K-Theory 8 (2011), no. 2, 223–239.
  • [QiL19] T. Qian and P. Li. Singular integrals and Fourier theory on Lipschitz boundaries. Science Press Beijing, Beijing; Springer, Singapore, 2019.
  • [Ran98] N. Randrianantoanina. Hilbert transform associated with finite maximal subdiagonal algebras. J. Austral. Math. Soc. Ser. A 65 (1998), no. 3, 388–404.
  • [Ran02] N. Randrianantoanina. Spectral subspaces and non-commutative Hilbert transforms. Colloq. Math. 91 (2002), no. 1, 9–27.
  • [Ren04] A. Rennie. Summability for nonunital spectral triples. KK-Theory 31 (2004), no. 1, 71–100.
  • [RoW01] D. Rolfsen ad B. Wiest. Free group automorphisms, invariant orderings and topological applications. Algebr. Geom. Topol. 1 (2001), 311–320.
  • [RLL00] M. Rørdam, F. Larsen and N. Laustsen. An introduction to KK-theory for CC^{*}-algebras. London Math. Soc. Stud. Texts, 49. Cambridge University Press, Cambridge, 2000.
  • [ScS22] H. Schulz-Baldes and T. Stoiber. Harmonic analysis in operator algebras and its applications to index theory and topological solid state systems. Math. Phys. Stud. Springer, Cham, 2022.
  • [ScS23] H. Schulz-Baldes and T. Stoiber. The generators of the K-groups of the sphere. Expo. Math. 41 (2023), no. 4, Paper No. 125519.
  • [Sud17] T. Sudo. The K-theory for the CC^{*}-algebras of nilpotent discrete groups by examples. Bull. Fac. Sci. Univ. Ryukyus (2017), no. 104, 1–40.
  • [SuZ23] F. Sukochev and D. Zanin. The Connes character formula for locally compact spectral triples. Astérisque (2023), no. 445.
  • [FXZ23] F. Sukochev, X. Xiong and D. Zanin. Asymptotics of singular values for quantised derivatives on noncommutative tori. J. Funct. Anal. 285 (2023), no. 5, Paper No. 110021.
  • [Tch01] P. Tchamitchian. The solution of Kato’s conjecture (after Auscher, Hofmann, Lacey, McIntosh and Tchamitchian). Journées “Équations aux Dérivées Partielles” (Plestin-les-Grèves, 2001), Exp. No. XIV, 14 pp., Univ. Nantes, Nantes, 2001.
  • [Uch78] A. Uchiyama. On the compactness of operators of Hankel type. Tohoku Math. J. (2) 30 (1978), no. 1, 163–171.
  • [Ver21] J. Verdera. Birth and life of the L2L^{2} boundedness of the Cauchy Integral on Lipschitz graph. Preprint, arXiv:2109.06690.
  • [Voi92] J. Voigt. On the convex compactness property for the strong operator topology. Note Mat. 12 (1992), 259–269.
  • [Wea96] N. Weaver. Lipschitz algebras and derivations of von Neumann algebras. J. Funct. Anal. 139 (1996), no. 2, 261–300.
  • [WeO93] N. E. Wegge-Olsen. K-theory and CC^{*}-algebras. A friendly approach. Oxford Sci. Publ. The Clarendon Press, Oxford University Press, New York, 1993.
  • [Wic20] B. D. Wick. Commutators, BMO, Hardy spaces and factorization: a survey. Real Anal. Exchange 45 (2020), no. 1, 1–28.
  • [Wir22] M. Wirth. The Differential Structure of Generators of GNS-symmetric Quantum Markov Semigroups. Preprint, arXiv:2207.09247.
  • [XXY18] X. Xiong, Q. Xu and Z. Yin. Sobolev, Besov and Triebel-Lizorkin spaces on quantum tori. Mem. Amer. Math. Soc. 252 (2018), no. 1203.

Cédric Arhancet
6 rue Didier Daurat, 81000 Albi, France
URL: https://sites.google.com/site/cedricarhancet
[email protected]
ORCID: 0000-0002-5179-6972