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

\NewEnviron

malign

(0.1) \BODY\displaystyle\begin{split}\BODY\end{split}

Topological Classification of Insulators:
I. Non-interacting Spectrally-Gapped One-Dimensional Systems

Jui-Hui Chung
Department of Applied Mathematics, Princeton University
Jacob Shapiro
Department of Mathematics, Princeton University
Abstract

We study non-interacting electrons in disordered one-dimensional materials which exhibit a spectral gap, in each of the ten Altland-Zirnbauer symmetry classes. We define an appropriate topology on the space of Hamiltonians so that the so-called strong topological invariants become complete invariants yielding the one-dimensional column of the Kitaev periodic table, but now derived without recourse to K-theory. We thus confirm the conjecture regarding a one-to-one correspondence between topological phases of gapped non-interacting 1D systems and the respective Abelian groups {0},,2,2\Set{0},\mathbb{Z},2\mathbb{Z},\mathbb{Z}_{2} in the spectral gap regime. The main tool we develop is an equivariant theory of homotopies of local unitaries and orthogonal projections. Moreover, we discuss an extension of the unitary theory to partial isometries, to provide a perspective towards the understanding of strongly-disordered, mobility-gapped materials.

1 Introduction

Topological insulators [HK10] are exotic materials which insulate in their bulk, but may be excellent conductors along their boundary. The quintessential example is Galium-Arsenic in two dimensions, at very low temperatures and strong perpendicular magnetic fields, which exhibits the celebrated integer quantum Hall effect (IQHE) [KDP80]. Beyond the aforementioned typical bulk-boundary behavior [Gra07], another defining feature of these materials is that they exhibit observables which are quantized and experimentally stable–a manifestation of macroscopic quantum mechanical effects. Mathematically this phenomenon suggests a global, topological description and indeed Nobel prizes have been awarded [THK16] for the association of the integer quantum Hall effect with the mathematical theory of algebraic topology, see e.g. [Tho+82, ASS83]. A decisive step was taken by Kitaev [Kit09] who devised a periodic Table˜1 of insulators organized by the Altland-Zirnbauer symmetry classes [AZ97] and patterned after K-theoretic Bott periodicity. The classification problem which is in present focus here enjoyed much attention recently in the mathematics literature, from various perspectives, see e.g. [FM13, DG15, Kub16, Thi16, BCR16, PS16, GS16, KK18, Kel19, AMZ20, BS20, BO21, AT22, GMP22].

Symmetry dimension
AZ Θ\hskip 4.2679pt\Theta\hskip 4.2679pt Ξ\hskip 4.2679pt\Xi\hskip 4.2679pt Π\hskip 4.2679pt\Pi\hskip 4.2679pt 11 22 33 44 55 66 77 88
A 0 0 0 0 \mathbb{Z} 0 \mathbb{Z} 0 \mathbb{Z} 0 \mathbb{Z}
AIII 0 0 11 \mathbb{Z} 0 \mathbb{Z} 0 \mathbb{Z} 0 \mathbb{Z} 0
AI 11 0 0 0 0 0 \mathbb{Z} 0 2\mathbb{Z}_{2} 2\mathbb{Z}_{2} \mathbb{Z}
BDI 11 11 11 \mathbb{Z} 0 0 0 \mathbb{Z} 0 2\mathbb{Z}_{2} 2\mathbb{Z}_{2}
D 0 11 0 2\mathbb{Z}_{2} \mathbb{Z} 0 0 0 \mathbb{Z} 0 2\mathbb{Z}_{2}
DIII 1-1 11 11 2\mathbb{Z}_{2} 2\mathbb{Z}_{2} \mathbb{Z} 0 0 0 \mathbb{Z} 0
AII 1-1 0 0 0 2\mathbb{Z}_{2} 2\mathbb{Z}_{2} \mathbb{Z} 0 0 0 \mathbb{Z}
CII 1-1 1-1 11 \mathbb{Z} 0 2\mathbb{Z}_{2} 2\mathbb{Z}_{2} \mathbb{Z} 0 0 0
C 0 1-1 0 0 \mathbb{Z} 0 2\mathbb{Z}_{2} 2\mathbb{Z}_{2} \mathbb{Z} 0 0
CI 11 1-1 11 0 0 \mathbb{Z} 0 2\mathbb{Z}_{2} 2\mathbb{Z}_{2} \mathbb{Z} 0
Table 1: The Kitaev periodic table. The entries stand for the respective K-theory groups in a given dimension and symmetry class. The present paper focuses on the one-dimensional column. See Section˜5 for explanation of Θ,Ξ,Π\Theta,\Xi,\Pi.

A first presentation of the association of quantum mechanics of insulators with algebraic topology would assume periodicity of the materials involved, which leads very naturally to the theory of equivariant vector bundles and their classification via K-theory, culminating in, e.g., [FM13]. However, while vector bundle theory is mathematically classical, a periodic model cannot describe strong-disorder, an important feature of topological insulators (see below). This has been recognized early on by Bellissard and collaborators [BvS94] who have laid important ground work in the 1990s to build bridges from physics into K-theory of C-star algebras and use Connes’ tools from non-commutative geometry to study what they refer to as the non-commutative Brillouin zone. And yet using K-theory bears a price: it allows homotopies to explore additional internal degrees of freedom, and it only studies relative phases. These two points mean the classification is more fuzzy than one would hope for (this point receives some attention in [DG15]). For this reason one might argue that K-theoretic classifications do not offer a one-to-one correspondence between topological phases of gapped systems and the respective Abelian K-theory groups. More severely, there does not seem to be a way to extend it to the strongly-disordered mobility gap regime–the description remains in the disordered spectral gap regime. Moreover, K-theory of C-star algebras with real or quaternionic structures is difficult to handle since (as far as we are aware) its dual, which is necessary to study index pairings, is not defined. These latter two points are somewhat addressed by Kasparov’s KK-theory, which is however vastly more complicated and (as far as we are aware) still cannot address the mobility gap problem.

Let us expand on the mobility gap regime briefly. The physical situation of materials being insulators is encoded mathematically by operators that have a certain gap. In the simplest scenario this is a spectral gap about the Fermi energy. But it turns out that when strong disorder is present (i.e., under Anderson localization) this spectral gap closes and the Fermi energy is immersed in an interval of localized states which cannot contribute to electric conductance, a situation referred to as the mobility gap regime [AG98]. These localized states are however essential in order to explain many important features of topological insulators, e.g., why plateaus emerge in the integer quantum Hall effect; see [EGS05, GS18, ST19, Sha20, BSS23] for further discussion of the mobility gap regime. In the spectral gap regime, the Fermi projection is a continuous function of the Hamiltonian and thus belongs to the C-star algebra generated by it. This makes the spectral gap regime amenable to analysis by K-theory of C-star algebras. On the other hand, in the mobility gap regime, the Fermi projection is merely a measurable function of the Hamiltonian.

It is mainly for the study of the mobility gap problem that it is important to be able to build alternative perspectives to the classification problem that do not rely on algebraic topology of classical manifolds (as in the periodic case) or on K-theory of C-star algebras (as in the disordered but spectrally gapped case), and this is the main point of the present paper: we present a first K-theoretic-free classification of disordered materials to our knowledge. Moreover, in our approach the question of “which topology to define on the space of ‘insulators’ ” becomes explicit and is brought to the foreground, since without it one cannot even start the analysis.

It remains unclear just what physical (better yet, experimentally relevant) role the choice of this topology bears, and it is also interesting to ask whether this choice is necessarily unique (we presume it is not). Be that as it may, since topological insulators are presumed useful for quantum computing [KKR06], where it is precisely the topological stability properties that lend themselves to be of great utility, it seems that exploring the foundations and boundaries of these stability properties could maybe help answer edge cases of quantum engineering problems.

In this paper we build the first step of this research program, which is the most straight-forward, namely, understanding non-interacting one-dimensional spectrally-gapped disordered systems via homotopies and without K-theory. This has the appeal that it is simpler–though this is a matter of taste–than the existing K-theoretic classifications, but also, that it allows us to start working on the next steps in the aforementioned program:

  1. 1.

    Higher dimensions in the spectral-gap, non-interacting case, and a more detailed study of higher dimensional locality (see Section˜8).

  2. 2.

    The strongly-disordered mobility gap regime (see Section˜7).

  3. 3.

    The interacting case (and within it the fractional quantum Hall effect).

  4. 4.

    Understanding interactions in the strongly-disordered regime, and hence also many-body localization (MBL).

It is mainly the second item which we feel is amenable to the methods developed here.

Let us briefly describe the mathematical novelty of this paper, to be presented in Sections˜3 and 4. Quantum mechanical Hamiltonians, beyond being self-adjoint, must obey a certain kind of locality constraint which is central in the present paper. Indeed it is that constraint which elevates the analysis from pure functional analysis into physics. This constraint roughly corresponds to the fact that there is no action at a distance. Geometrically this can be understood as a non-commutative analog of a regularity constraint on symbols, since, if our systems were translation invariant it would correspond to continuity of the symbol via a Riemann-Lebesgue lemma. Hence we are concerned with spaces of local operators. Under the various symmetry constraints these operators break down into two main classes depending on the presence or absence of a so-called chiral symmetry: unitaries or self-adjoint projections. These two broad categories are then broken into five additional ones: complex, real, quaternionic, and so-called \star-real or \star-quaternionic (see Section˜2). Hence all together we find ten possible classes. Let us consider, then, the simplest case: that of complex unitaries. Without the locality constraint, it is a result that goes back to Kuiper [Kui65] (see Theorem˜B.1 below) that the set of unitaries on a separable Hilbert space is path-connected. Indeed, a path from any unitary UU to 𝟙\mathds{1} is given by

(1.1) [0,1]texp(i(1t)(ilog(U)))\displaystyle[0,1]\ni t\mapsto\exp\left(\operatorname{i}(1-t)\left(-\operatorname{i}\log\left(U\right)\right)\right)

where ilog(U)-\operatorname{i}\log\left(U\right) is a self-adjoint operator to be understood via the bounded measurable functional calculus of normal operators. In contrast to Kuiper’s situation, the space of local unitaries turns out to be very much disconnected: the components are indexed by a non-commutative analog of the winding number, which under the assumption of translation invariance indeed collapses to the classical winding number (this is the Krein-Widom-Devinatz theorem [Dou98, pp. 185]). The winding number requires the continuity of the map to be meaningful, which is analogous to the present locality constraint. The main issue to be dealt with is, then: given two local unitaries U,VU,V of the same index, construct a continuous local path between them, or equivalently, given a local unitary of zero index, connect it locally to 𝟙\mathds{1}. It is a theorem that if a local unitary has non-zero index then its spectrum is the whole 𝕊1\mathbb{S}^{1} [ABJ20]. Naively one might expect that unitaries of zero index always have a spectral gap on 𝕊1\mathbb{S}^{1} and hence the above logarithm may actually be interpreted via the holomorphic functional calculus, in which case it preserves locality (this is the Combes-Thomas estimate for unitaries, see e.g. [HJS09, ST19]). This is unfortunately false: take as a counter-example any continuous map 𝕊1𝕊1\mathbb{S}^{1}\to\mathbb{S}^{1} which has zero winding number but whose range is 𝕊1\mathbb{S}^{1}. Its Fourier series will correspond to a local unitary of zero index which has σ(U)=𝕊1\sigma(U)=\mathbb{S}^{1}. The solution is then to factorize U=ABU=AB where A,BA,B are two local unitaries, one of which has a gap and the other diagonal in a left-right decomposition of the Hilbert space, and is hence amenable to a (not necessarily local) usual Kuiper path on each side of space separately. The homotopies of local complex unitaries were first studied in [CHO82], although there a different proof was presented. The non-complex local unitary homotopies are, to our knowledge, new. For self-adjoint projections the local homotopies are somewhat different; to this end we make equivariant extensions of the work of [ACL15]. It turns out that in the complex case, all self-adjoint local projections of a certain non-trivial class are path-connected.

In two of the symmetry classes, the index is 2\mathbb{Z}_{2}-valued, corresponding to the Atiyah-Singer skew-adjoint Fredholm index [AS69], see Appendix˜A for an introduction. For these symmetry classes, the analysis becomes more complicated due to the absence of a logarithmic law for the 2\mathbb{Z}_{2} index, leading us to connect directly two arbitrary operators of odd index. The application of Atiyah and Singer’s skew-adjoint Fredholm index in the context of topological insulators was pioneered in [Sch15] but then studied also in [KK16, Fon+20, BSS23].

In regards to existing literature, almost exclusively, classification results of topological insulators rely on K-theory and it is in this sense that they do not provide a complete homotopy classification. Of the ones listed in the first paragraph above, we mention the paper by Thiang [Thi16] who provides a K-theoretic classification of disordered spectrally gapped systems in all dimensions. On a more pedestrian note, if one assumes translation invariance, the classification problem is of course classical and reduces to studying homotopies of continuous maps 𝕋dGrk(N)\mathbb{T}^{d}\to\mathrm{Gr}_{k}(\mathbb{C}^{N}) under various symmetry constraints where Grk(N)\mathrm{Gr}_{k}(\mathbb{C}^{N}) is the Grassmannian: the space of kk-dimensional subspaces within N\mathbb{C}^{N}. This classification is in fact known to “contradict” Table˜1 due to: (1) low NN problems, and (2) the existence of weak topological invariants. These are, roughly speaking, indices which do not explore all dd dimensions of real space and are not stable under strong disorder. Recently Avron and Turner [AT22] presented a full classification of these translation invariant systems in the special case d=k=1=N/2d=k=1=N/2.

This paper is organized as follows. In Section˜2 we present the abstract mathematical setting of odd-dimensional locality, symmetries and the associated indices. This section is mainly intended to set up the terminology and notation for Sections˜3 and 4 in which we calculate π0\pi_{0} of various symmetry-constrained local unitaries and self-adjoint projections. We make use of this theory in Section˜5 by connecting it to the problem of classifying bulk one-dimensional spectrally-gapped insulators. Within this section, we single out Section˜5.6 where operators with the more common form of exponential locality are studied using an entirely separate scheme. After making some brief remarks about edge systems in Section˜6, we present a negative result about the classification in the mobility gap result in Section˜7 and conclude in Section˜8 with a few words about the classification problem in higher dimensions. We shall argue there that even though in some sense one may wish to draw conclusions from our work on the classification problem in all odd dimensions, the notion of locality we employ here and which makes sense in one-dimension, is rather unsatisfactory in higher dimensions, which warrants that not only the even-dimensional but also the odd-dimensional problem be revisited in future work.

Notations and conventions

  • π0\pi_{0} is the path-components functor acting on the category of topological spaces.

  • We use |A|2AA|A|^{2}\equiv A^{\ast}A and pol(A)\operatorname{pol}(A) for the polar part in the polar decomposition A=pol(A)|A|A=\operatorname{pol}(A)|A|, made unique by the convention that kerU=kerA\ker U=\ker A.

  • \mathcal{H} is a separable Hilbert space, \mathcal{B} is its Banach algebra of bounded linear operators, and 𝒰,𝒢,,𝒦\mathcal{U},\mathcal{G},\mathcal{F},\mathcal{K} are the subspaces of unitary, invertible, Fredholm, and compact operators respectively. We shall also use the space 𝒫\mathcal{P} of self-adjoint (orthogonal) projections and (the equivalent) 𝒮\mathcal{S}, the space of self-adjoint unitary operators. Sometimes we also use sa\mathcal{F}^{\mathrm{sa}} for the space of self-adjoint Fredholm operators.

  • For us AA is idempotent iff A2=AA^{2}=A and AA is a self-adjoint (orthogonal) projection iff A2=A=AA^{2}=A=A^{\ast}. We generally try to avoid the term “projection” by itself since some authors use it for idempotent and others for “self-adjoint projection”.

  • \mathcal{L} is the C-star algebra of local operators, those operator having a compact commutator with a fixed projection Λ\Lambda.

  • CC is a real structure (anti-unitary) which squares to +𝟙+\mathds{1} and JJ is a quaternionic structure (anti-unitary) which squares to 𝟙-\mathds{1}.

  • By the word “essentially” we generally mean that an algebraic condition holds up to compacts, i.e., in the Calkin algebra. With this, we have essentially unitary operators (𝟙|A|2,𝟙|A|2𝒦\mathds{1}-|A|^{2},\mathds{1}-|A^{\ast}|^{2}\in\mathcal{K}), essentially projections (AA,AA2𝒦A-A^{\ast},A-A^{2}\in\mathcal{K}), etc.

  • We use 𝕊1\mathbb{S}^{1} for the unit circle and 𝔻\mathbb{D} for the open unit disc, both understood as subsets of \mathbb{C}.

2 Abstract locality, indices and symmetry constraints

In this section, \mathcal{H} is some fixed abstract separable Hilbert space.

Definition 2.1 (non-trivial projections).

We call a self-adjoint projection P𝒫P\in\mathcal{P} non-trivial iff its range and kernel are both infinite dimensional.

Essential projections are classical objects which go back to [Cal41]: AA\in\mathcal{B} is called essentially a projection iff AA,AA2𝒦A-A^{\ast},A-A^{2}\in\mathcal{K}. Actually this implies, via Lemma˜B.8, that there exists some P𝒫P\in\mathcal{P} such that AP𝒦A-P\in\mathcal{K}. Less common is the notion of essentially non-trivial projections:

Definition 2.2 (essentially a non-trivial projection).

We call a bounded linear operator AA\in\mathcal{B} “essentially a non-trivial projection” iff AA is essentially a projection and {0,1}σess(A)\Set{0,1}\subseteq\sigma_{\mathrm{ess}}(A).

It will be useful to have another criterion for essentially non-trivial projections:

Lemma 2.3.

AA\in\mathcal{B} is essentially a non-trivial projection iff there exists a non-trivial projection PP such that AP𝒦A-P\in\mathcal{K}. Furthermore, if BB\in\mathcal{B} is essentially a projection and ABA-B is sufficiently small in norm or compact, then BB is also essentially a non-trivial projection.

Proof.

If there exists a non-trivial projection PP such that AP𝒦A-P\in\mathcal{K}, write AP=KA-P=K, then AA=KK𝒦A-A^{*}=K-K^{*}\in\mathcal{K} and AA2=P+K(P+K)2=KPKKPK2𝒦A-A^{2}=P+K-(P+K)^{2}=K-PK-KP-K^{2}\in\mathcal{K} and hence AA is essentially a projection. In particular, {0,1}σess(P)=σess(P+K)=σess(A)\Set{0,1}\subseteq\sigma_{\mathrm{ess}}(P)=\sigma_{\mathrm{ess}}(P+K)=\sigma_{\mathrm{ess}}(A). Thus AA is essentially a non-trivial projection.

Conversely, if AA is essentially a projection, using Lemma˜B.8, there exists a self-adjoint projection PP such that AP𝒦A-P\in\mathcal{K}. In particular, {0,1}σess(A)=σess(P+K)=σess(P)\Set{0,1}\subseteq\sigma_{\mathrm{ess}}(A)=\sigma_{\mathrm{ess}}(P+K)=\sigma_{\mathrm{ess}}(P), so the kernel and image of PP are infinite and thus PP is non-trivial.

Now, if AB𝒦A-B\in\mathcal{K} the statement is trivial. Moreover, we note that if AA is essentially a projection, then, it is non-trivial iff A,𝟙AA,\mathds{1}-A are not Fredholm (by the Fredholm definition of the essential spectrum). Hence, if AB\left\lVert A-B\right\rVert is sufficiently small, it can’t be that BB is Fredholm whereas AA is not, since the Fredholms are open, and same with (𝟙A)(𝟙B)(\mathds{1}-A)-(\mathds{1}-B). Thus {0,1}σess(B)\Set{0,1}\in\sigma_{\mathrm{ess}}(B). ∎

Remark 2.4.

Given a projection P𝒫P\in\mathcal{P}, 𝟙2P\mathds{1}-2P is a self-adjoint unitary, so that the space 𝒫\mathcal{P} is identified with

𝒮{U𝒰|U=U},\displaystyle\mathcal{S}\equiv\Set{U\in\mathcal{U}}{U=U^{\ast}}\,,

the space of self-adjoint unitaries, and all the notions discussed above of non-triviality of projections carry over to self-adjoint unitaries. We shall refer to both spaces interchangeably, and the classification of local self-adjoint projections or local self-adjoint unitaries is the same.

Definition 2.5 (Λ\Lambda-local operators).

For a fixed non-trivial self-adjoint projection Λ\Lambda, an operator AA\in\mathcal{B} is termed Λ\Lambda-local iff it essentially commutes with Λ\Lambda, i.e.,

(2.1) [Λ,A]ΛAAΛ𝒦.\displaystyle[\Lambda,A]\equiv\Lambda A-A\Lambda\in\mathcal{K}\,.

The space of all local operators is denoted by Λ\mathcal{L}_{\Lambda}. Clearly if a projection is trivial, the condition is vacuous, and hence the restriction. Sometimes we use the phrase hyper-local if [A,Λ]=0[A,\Lambda]=0.

Unless otherwise specified (mainly relevant in Section˜7) we shall always use the subspace topology induced by the operator norm topology on \mathcal{B} unless otherwise specified. With respect to this topology, we use π0\pi_{0} as the path-components functor.

For most of what follows, we shall not have occasion to consider different Λ\Lambda’s for locality, and so, let us fix once and for all one self-adjoint projection Λ\Lambda and omit this choice entirely from the notation. If a space 𝒜\mathcal{A} carries the superscript \mathcal{L} we mean by it the intersection:

(2.2) 𝒜𝒜\displaystyle\mathcal{A}^{\mathcal{L}}\equiv\mathcal{A}\cap\mathcal{L}

and the prefix 𝒮\mathcal{S} means the subset of self-adjoint operators within 𝒜\mathcal{A}.

Lemma 2.6.

\mathcal{L} is a C-star algebra with respect to the operator norm and adjoint inherited from \mathcal{B}.

Proof.

The only thing to verify is the compact commutator condition is closed. However, the norm limit of compact operators is compact, and hence the statement follows. ∎

Remark 2.7.

To the extent that commutators may be considered as non-commutative derivatives, locality may be thought of as a certain regularity condition analogous to differentiability. This is essentially Bellissard et al’s non-commutative Sobolev spaces [BvS94].

Lemma 2.8.

The continuous functional calculus on normal operators maps \mathcal{L} to \mathcal{L}.

Proof.

Let AA\in\mathcal{L} be normal and f:f:\mathbb{C}\to\mathbb{C} continuous. Since AA is bounded, its spectrum is restricted to some compact set SS\subseteq\mathbb{C} and hence we may assume WLOG that ff has support SS. Let now {pk:}k\Set{p_{k}:\mathbb{C}\to\mathbb{C}}_{k} be a sequence polynomials converging uniformly to ff on SS. Then pk(A)f(A)p_{k}(A)\to f(A) in operator norm, and hence, since each [pk(A),Λ][p_{k}(A),\Lambda] is compact (recall \mathcal{L} is a C-star algebra) its norm limit is too. ∎

We note in passing that for holomorphic functions (which may be desired when dealing with non-normal operators) this can be deduced by a Combes-Thomas type argument: the resolvent of a local operator is clearly local by [Λ,(Az𝟙)1]=(Az𝟙)1[Λ,A](Az𝟙)1[\Lambda,(A-z\mathds{1})^{-1}]=-(A-z\mathds{1})^{-1}[\Lambda,A](A-z\mathds{1})^{-1}.

We now define the so-called “super” operator Λ:\mathbb{\Lambda}:\mathcal{B}\to\mathcal{B} given by

(2.3) AΛAΛAΛ+𝟙Λ.\displaystyle\mathcal{B}\ni A\mapsto\mathbb{\Lambda}A\equiv\Lambda A\Lambda+\mathds{1}-\Lambda\,.

With it we may define an index for local unitaries as follows.

Lemma 2.9 (Fredholm property of local unitaries).

The image of 𝒰\mathcal{U}^{\mathcal{L}} under Λ\mathbb{\Lambda} is Fredholm, i.e.,

Λ(𝒰).\displaystyle\mathbb{\Lambda}(\mathcal{U}^{\mathcal{L}})\subseteq\mathcal{F}\,.
Proof.

Let U𝒰U\in\mathcal{U}^{\mathcal{L}}. Then using Atkinson’s theorem [BB89] it suffices to exhibit (ΛU)(\mathbb{\Lambda}U)^{\ast} as the parametrix of ΛU\mathbb{\Lambda}U, and to that end, we note that

𝟙(ΛU)(ΛU)=𝟙(ΛU)(ΛU)=Λ(𝟙UΛU)Λ=ΛUΛUΛ=ΛUΛ[U,Λ].\displaystyle\mathds{1}-(\mathbb{\Lambda}U)^{\ast}(\mathbb{\Lambda}U)=\mathds{1}-(\mathbb{\Lambda}U^{\ast})(\mathbb{\Lambda}U)=\Lambda(\mathds{1}-U^{\ast}\Lambda U)\Lambda=\Lambda U^{\ast}\Lambda^{\perp}U\Lambda=\Lambda U^{\ast}\Lambda^{\perp}[U,\Lambda]\,.

Now, since UU\in\mathcal{L} this last commutator is compact, and so by the ideal property of 𝒦\mathcal{K} the entire expression. ∎

It thus makes sense to define indΛ:𝒰\operatorname{ind}_{\Lambda}:\mathcal{U}^{\mathcal{L}}\to\mathbb{Z} via

(2.4) indΛ(U):=ind(ΛU)dimkerΛUdimkerΛU.\displaystyle\operatorname{ind}_{\Lambda}(U):=\operatorname{ind}(\mathbb{\Lambda}U)\equiv\dim\ker\mathbb{\Lambda}U-\dim\ker\mathbb{\Lambda}U^{\ast}\,.

This index reduces to the winding number, if the unitary happens to be a Toeplitz operator on 2()\ell^{2}(\mathbb{Z}) (this statement is the aforementioned Krein-Widom-Devinatz theorem [Dou98, pp. 185], sometimes also referred to as the Krein-Gohberg theorem). It is comforting to know that this index inherits the logarithm law from the Fredholm index:

Lemma 2.10 (logarithmic law).

If U,V𝒰U,V\in\mathcal{U}^{\mathcal{L}} then

(2.5) indΛU+indΛV=indΛ(UV)\displaystyle\operatorname{ind}_{\Lambda}U+\operatorname{ind}_{\Lambda}V=\operatorname{ind}_{\Lambda}(UV)
Proof.

Using the logarithmic law of ind:\operatorname{ind}:\mathcal{F}\to\mathbb{Z} [BB89], it remains to show Λ(UV)(ΛU)ΛV\mathbb{\Lambda}(UV)-(\mathbb{\Lambda}U)\mathbb{\Lambda}V is compact. This follows from

Λ(UV)(ΛU)ΛV=ΛUVΛΛUΛVΛ=ΛUΛVΛ=[Λ,U]𝒦ΛVΛ.\displaystyle\mathbb{\Lambda}(UV)-(\mathbb{\Lambda}U)\mathbb{\Lambda}V=\Lambda UV\Lambda-\Lambda U\Lambda V\Lambda=\Lambda U\Lambda^{\perp}V\Lambda=\underbrace{[\Lambda,U]}_{\in\mathcal{K}}\Lambda^{\perp}V\Lambda\,.

We now turn to symmetry constraints. Let C,J:C,J:\mathcal{H}\to\mathcal{H} be two fixed anti-unitary operators on \mathcal{H} such that

C2=𝟙,J2=𝟙.\displaystyle C^{2}=\mathds{1}\,,\qquad\,J^{2}=-\mathds{1}.

As such, CC and JJ define real and quaternionic structures respectively on \mathcal{H}: CC should be understood as complex conjugation and JJ as the jjth quaternionic basis vector, so that 𝟙,i𝟙,J\mathds{1},\operatorname{i}\mathds{1},J and iJ\operatorname{i}J build the quaternionic basis vectors [Bae12]. It is thus natural to consider the subspace of real and quaternionic bounded operators, those which respect that structure:

(2.6) :={A|AC=CA},:={A|AJ=JA}.\displaystyle\mathcal{B}_{\mathbb{R}}:=\Set{A\in\mathcal{B}}{AC=CA}\,,\qquad\mathcal{B}_{\mathbb{H}}:=\Set{A\in\mathcal{B}}{AJ=JA}\,.

We note that in the latter case, unitary operators UU obeying [U,J]=0[U,J]=0 may also be understood as Hermitian-symplectic operators (discussed e.g. in [Sha21, (3.7)]) with respect to the symplectic bi-linear form given by ,J\langle\cdot,J\cdot\rangle, since then one has UJU=JU^{\ast}JU=J and hence the bi linear form ,J\langle\cdot,J\cdot\rangle is preserved by such UU.

We shall also need the following somewhat more exotic symmetry constraints. For lack of better terminology, we call them \star-real and \star-quaternionic operators:

(2.7) :={A|AC=CA},:={A|AJ=JA}.\displaystyle\mathcal{B}_{\star\mathbb{R}}:=\Set{A\in\mathcal{B}}{AC=CA^{\ast}}\,,\qquad\mathcal{B}_{\star\mathbb{H}}:=\Set{A\in\mathcal{B}}{AJ=JA^{\ast}}\,.

In [Fon+20] we used the terminology JJ-odd for the same constraint (only \mathcal{B}_{\star\mathbb{H}} was used there), but in the current abstract mathematical setting it is more natural to use the real and quaternionic structures. We caution the reader that our naming is not standard, e.g., in [GP06] the name CC-symmetric was used for \mathcal{B}_{\star\mathbb{R}}.

The following purely imaginary classes are not independent of the ones presented so far, but we introduce them separately nonetheless for notational simplicity:

(2.8) i:={A|AC=CA},i:={A|AJ=JA}.\displaystyle\mathcal{B}_{\operatorname{i}\mathbb{R}}:=\Set{A\in\mathcal{B}}{AC=-CA}\,,\qquad\mathcal{B}_{\operatorname{i}\mathbb{H}}:=\Set{A\in\mathcal{B}}{AJ=-JA}\,.

They may be obtained as i,i\operatorname{i}\mathcal{B}_{\mathbb{R}},\operatorname{i}\mathcal{B}_{\mathbb{H}} respectively.

We shall see below in Section˜5 that these combinations build together all the necessary Altland-Zirnbauer symmetry classes (the ten fold way) which appear in Table˜1.

Assumption 2.11 (real and quaternionic structures are hyper-local).

We shall assume that C,JC,J are chosen so that

(2.9) [J,Λ]=[C,Λ]=0.\displaystyle[J,\Lambda]=[C,\Lambda]=0\,.

This can probably be weakened to from zero to compact, but we do not need this generalization.

It is then clear that, for 𝔽=,\mathbb{F}=\mathbb{R},\mathbb{H}, restricting indΛ\operatorname{ind}_{\Lambda} to 𝒰𝔽\mathcal{U}_{\star\mathbb{F}}^{\mathcal{L}}, we get the constant zero map. Indeed, this is immediate from the fact C,JC,J are bijections and the logarithmic rule Lemma˜2.10. The same is true within 𝒮𝔽\mathcal{S}^{\mathcal{L}}_{\mathbb{F}} for any 𝔽\mathbb{F} by self-adjointness. Be that as it may, Atiyah and Singer recognized that another index, a 2\mathbb{Z}_{2} index, may sometimes be defined (see Appendix˜A below):

(2.10) indΛ,2(U):=ind2ΛU(dimkerΛUmod2)2\displaystyle\operatorname{ind}_{\Lambda,2}(U):=\operatorname{ind}_{2}\mathbb{\Lambda}U\equiv\left(\dim\ker\mathbb{\Lambda}U\mod 2\right)\in\mathbb{Z}_{2}

where ind2:2\operatorname{ind}_{2}:\mathcal{F}\to\mathbb{Z}_{2} is the Atiyah-Singer 2\mathbb{Z}_{2} Fredholm index. As discussed in Theorems˜A.2 and A.3, this index is norm continuous as a map with domain 𝒰\mathcal{U}_{\star\mathbb{H}}^{\mathcal{L}} or 𝒮i\mathcal{S}^{\mathcal{L}}_{\operatorname{i}\mathbb{R}} respectively.

3 Equivariant classification of local unitaries

In this section we shall study π0(𝒰𝔽)\pi_{0}(\mathcal{U}_{\mathbb{F}}^{\mathcal{L}}) where 𝔽\mathbb{F} is either \mathbb{C} (in which case this is just the space of local unitaries) or 𝔽\mathbb{F} is one of the four symmetries discussed above: ,,,\mathbb{R},\mathbb{H},\star\mathbb{R},\star\mathbb{H}. We group our theorems together based on method of proof. The results are summarized in Table˜2.

We start with the main classification statement:

Theorem 3.1 (Classification of ,\mathbb{R},\mathbb{C} and \mathbb{H} local unitaries).

For 𝔽=,\mathbb{F}=\mathbb{R},\mathbb{C} the map indΛ:U𝔽\operatorname{ind}_{\Lambda}:U_{\mathbb{F}}^{\mathcal{L}}\to\mathbb{Z} is norm continuous and ascends to a bijection

(3.1) indΛ:π0(𝒰𝔽)\displaystyle\operatorname{ind}_{\Lambda}:\pi_{0}(\mathcal{U}_{\mathbb{F}}^{\mathcal{L}})\xrightarrow{\sim}\mathbb{Z}

and analogously for the quaternionic class, we have the bijection

(3.2) indΛ:π0(𝒰)2.\displaystyle\operatorname{ind}_{\Lambda}:\pi_{0}(\mathcal{U}_{\mathbb{H}}^{\mathcal{L}})\xrightarrow{\sim}2\mathbb{Z}\,.

This theorem should be compared with Kuiper’s theorem (π0(𝒰){0}\pi_{0}(\mathcal{U})\cong\Set{0}, see Theorem˜B.1) and the Atiyah-Jänich theorem (ind:π0()\operatorname{ind}:\pi_{0}(\mathcal{F})\cong\mathbb{Z}) [BB89]. Strictly speaking, when 𝔽=\mathbb{F}=\mathbb{C}, it is not new: it may be deduced from the results of [CHO82], where the criterion of locality as a compact commutator is replaced by the commutator belonging to a more general ideal. We shall present a different proof, which also covers the cases 𝔽=,\mathbb{F}=\mathbb{R},\mathbb{H} (which as far as we are aware has not appeared previously). We also became aware that [Gei22] contains ideas of similar spirit.

Next, we have the nullhomotopic result:

Theorem 3.2 (Classification of \star\mathbb{R}-local unitaries).

The space of \star\mathbb{R}-local unitaries is null-homotopic:

(3.3) π0(𝒰){0}.\displaystyle\pi_{0}(\mathcal{U}_{\star\mathbb{R}}^{\mathcal{L}})\cong\Set{0}\,.

Finally, there is the 2\mathbb{Z}_{2} classification:

Theorem 3.3 (Classification of \star\mathbb{H}-local unitaries).

The space of \star\mathbb{H}-local unitaries has two path-components. The map indΛ,2:𝒰2\operatorname{ind}_{\Lambda,2}:\mathcal{U}_{\star\mathbb{H}}^{\mathcal{L}}\to\mathbb{Z}_{2} is norm continuous and ascends to a bijection

(3.4) indΛ,2:π0(𝒰)2.\displaystyle\operatorname{ind}_{\Lambda,2}:\pi_{0}(\mathcal{U}_{\star\mathbb{H}}^{\mathcal{L}})\xrightarrow{\sim}\mathbb{Z}_{2}\,.
π0(𝒰)\pi_{0}(\mathcal{U}^{\mathcal{L}}) \mathbb{Z}
π0(𝒰)\pi_{0}(\mathcal{U}_{\mathbb{R}}^{\mathcal{L}}) \mathbb{Z}
π0(𝒰)\pi_{0}(\mathcal{U}_{\mathbb{H}}^{\mathcal{L}}) 22\mathbb{Z}
π0(𝒰)\pi_{0}(\mathcal{U}_{\star\mathbb{R}}^{\mathcal{L}}) {0}\Set{0}
π0(𝒰)\pi_{0}(\mathcal{U}_{\star\mathbb{H}}^{\mathcal{L}}) 2\mathbb{Z}_{2}
Table 2: The classification of equivariant local unitaries.

The main technical tool to be used in Theorems˜3.1, 3.3 and 3.2 is a factorization principle, which we present and prove before tending to the proofs of the main theorems.

Lemma 3.4 (factorization of local unitaries).

For any 𝔽{,,}{\mathbb{F}\in\Set{\mathbb{C},\mathbb{R},\mathbb{H}}}, let U𝒰𝔽U\in\mathcal{U}_{\mathbb{F}}^{\mathcal{L}} such that

indΛU=0.\displaystyle\operatorname{ind}_{\Lambda}U=0\,.

Then there exist two unitaries A,B𝒰𝔽A,B\in\mathcal{U}_{\mathbb{F}}^{\mathcal{L}} such that 𝟙A𝒦\mathds{1}-A\in\mathcal{K}, [Λ,B]=0[\Lambda,B]=0 and such that

(3.5) U=AB.\displaystyle U=AB\,.
Proof.

Let 𝔽{,,}\mathbb{F}\in\Set{\mathbb{C},\mathbb{R},\mathbb{H}}, and F{𝟙,C,J}F\in\Set{\mathds{1},C,J} accordingly. Let us decompose UU in =(imΛ)imΛ\mathcal{H}=(\operatorname{im}\Lambda)^{\perp}\oplus\operatorname{im}\Lambda as

(3.6) U=[ULLULRURLURR].\displaystyle U=\begin{bmatrix}U_{LL}&U_{LR}\\ U_{RL}&U_{RR}\end{bmatrix}\,.

U𝒰U\in\mathcal{U}^{\mathcal{L}} implies both ULR,URL𝒦U_{LR},U_{RL}\in\mathcal{K}, and UiiU_{ii} is essentially unitary: 𝟙|Uii|2,𝟙|Uii|2𝒦\mathds{1}-|U_{ii}|^{2},\mathds{1}-|U_{ii}^{\ast}|^{2}\in\mathcal{K} (see [Mur90]). By Assumption˜2.11, the subspaces imΛ,(imΛ)\operatorname{im}\Lambda,(\operatorname{im}\Lambda)^{\perp} are invariant under the action of 𝔽\mathbb{F}-structures, i.e., FF is diagonal in this grading and (hopefully without confusion) we do not give each block a separate symbol. Thus UF=FUUF=FU implies UiiF=FUiiU_{ii}F=FU_{ii} for i{L,R}i\in\Set{L,R}. Now indΛU=0\operatorname{ind}_{\Lambda}U=0 is equivalent to indURR=0\operatorname{ind}U_{RR}=0. Using indAB=indA+indB\operatorname{ind}A\oplus B=\operatorname{ind}A+\operatorname{ind}B, the fact that UU is unitary, and that ind\operatorname{ind} is invariant under compact perturbations, we have

0=indU=indULLURR=indULL+indURR=indULL\displaystyle 0=\operatorname{ind}U=\operatorname{ind}U_{LL}\oplus U_{RR}=\operatorname{ind}U_{LL}+\operatorname{ind}U_{RR}=\operatorname{ind}U_{LL}

so that indULL=0\operatorname{ind}U_{LL}=0 as well. Hence UiiU_{ii} is an essential unitary Fredholm operator of zero index, so that applying Lemma˜B.4 below on UiiU_{ii} we obtain some Bii𝒰𝔽B_{ii}\in\mathcal{U}_{\mathbb{F}} which differs from UiiU_{ii} by a compact; we point out BiiB_{ii} is a unitary on one of the smaller spaces imΛ\operatorname{im}\Lambda,imΛ\operatorname{im}\Lambda^{\perp}.

Let B=BLLBRRB=B_{LL}\oplus B_{RR}. Define A=UBA=UB^{*} from which A𝒰𝔽A\in\mathcal{U}_{\mathbb{F}}^{\mathcal{L}} follows. To see that 𝟙A𝒦\mathds{1}-A\in\mathcal{K}, write Uii=Bii+KiiU_{ii}=B_{ii}+K_{ii} for some Kii𝒦K_{ii}\in\mathcal{K}. We have

(3.7) A=[BLL+KLLULRURLBRR+KRR][BLL00BRR]=[𝟙+KLLBLLULRBRRURLBLL𝟙+KRRBRR].\displaystyle A=\begin{bmatrix}B_{LL}+K_{LL}&U_{LR}\\ U_{RL}&B_{RR}+K_{RR}\end{bmatrix}\begin{bmatrix}B_{LL}^{*}&0\\ 0&B_{RR}^{*}\end{bmatrix}=\begin{bmatrix}\mathds{1}+K_{LL}B_{LL}^{*}&U_{LR}B^{*}_{RR}\\ U_{RL}B^{*}_{LL}&\mathds{1}+K_{RR}B^{*}_{RR}\end{bmatrix}\,.

For the \star classes, we need an adjusted factorization statement:

Lemma 3.5 (factorization of local \star-unitaries).

For any 𝔽{,}{\mathbb{F}\in\Set{\star\mathbb{R},\star\mathbb{H}}} and F{C,J}F\in\Set{C,J} accordingly, let U𝒰𝔽U\in\mathcal{U}_{\mathbb{F}}^{\mathcal{L}}. If 𝔽=\mathbb{F}=\star\mathbb{H}, we furthermore assume

indΛ,2U=0.\displaystyle\operatorname{ind}_{\Lambda,2}U=0\,.

Then there exist two unitaries A,BA,B with

U=AB\displaystyle U=AB

such that B𝒰𝔽B\in\mathcal{U}^{\mathcal{L}}_{\mathbb{F}} with [Λ,B]=0[\Lambda,B]=0, and such that AA is local, has 𝟙A𝒦\mathds{1}-A\in\mathcal{K}, and is furthermore 𝔽\star\mathbb{F} with respect to F~:=BF\widetilde{F}:=BF (i.e., A(BF)=(BF)AA(BF)=(BF)A^{*}).

Proof.

Let us decompose UU in =(imΛ)imΛ\mathcal{H}=(\operatorname{im}\Lambda)^{\perp}\oplus\operatorname{im}\Lambda as in eq.˜3.6 with the properties of UijU_{ij} listed there. Using Lemma˜B.6 below, for i=L,Ri=L,R, we have Bii𝒰𝔽B_{ii}\in\mathcal{U}_{\mathbb{F}} whose difference from UiiU_{ii} is compact. This is justified because indΛ,2Uii=0\operatorname{ind}_{\Lambda,2}U_{ii}=0. Indeed, by hypothesis, this holds for i=Ri=R, and since UU is unitary, ind2U=0\operatorname{ind}_{2}U=0. Now, using ˜A.4, we conclude that ind2ULLURR=0\operatorname{ind}_{2}U_{LL}\oplus U_{RR}=0. So ind2ULL=0\operatorname{ind}_{2}U_{LL}=0 too.

Let B=BLLBRR𝒰FB=B_{LL}\oplus B_{RR}\in\mathcal{U}_{F}^{\mathcal{L}} and A=UB𝒰A=UB^{*}\in\mathcal{U}^{\mathcal{L}}. We have 𝟙A𝒦\mathds{1}-A\in\mathcal{K} from eq.˜3.7. However, now 𝔽\mathcal{B}_{\mathbb{F}} is not an algebra. Nonetheless, observe that (BF)2=BFBF=FBBF=F2(BF)^{2}=BFBF=FB^{*}BF=F^{2} so that F~:=BF\widetilde{F}:=BF (instead of FF) defines an 𝔽\mathbb{F}-structure with which AA is \star-real or quaternionic:

A(BF)=UF=FU=FBA=(BF)A.\displaystyle A(BF)=UF=FU^{*}=FB^{*}A^{*}=(BF)A^{*}\,.

Using the factorization lemmas, we may tend to our three theorems. In regards to the continuity of ind(2),Λind(2)Λ\operatorname{ind}_{(2),\Lambda}\equiv\operatorname{ind}_{(2)}\circ\mathbb{\Lambda}, it is a consequence of the norm continuity of ind(2):(2)\operatorname{ind}_{(2)}:\mathcal{F}\to\mathbb{Z}_{(2)} and the trivial fact that Λ:\mathbb{\Lambda}:\mathcal{B}\to\mathcal{B} is continuous. This statement is true regardless of 𝔽\mathbb{F}.

So we merely need to show surjectivity (when applicable), well-definedness (when applicable) and injectivity.

Proof of Theorem˜3.1.

We start with surjectivity for 𝔽=\mathbb{F}=\mathbb{C}. Since Λ\Lambda is non-trivial, there is an ONB for \mathcal{H}, {φn}n\Set{\varphi_{n}}_{n\in\mathbb{Z}}, such that {φn}n0\Set{\varphi_{n}}_{n\leq 0} spans kerΛ\ker\Lambda and {φn}n>0\Set{\varphi_{n}}_{n>0} spans imΛ\operatorname{im}\Lambda. Define a unitary operator RR on \mathcal{H} via

Rφn:=φn+1(n).\displaystyle R\varphi_{n}:=\varphi_{n+1}\qquad(n\in\mathbb{Z})\,.

Since φn,[R,Λ]φm=(χ(n)χ(m))δn,m+1\langle\varphi_{n},[R,\Lambda]\varphi_{m}\rangle=\left(\chi_{\mathbb{N}}(n)-\chi_{\mathbb{N}}(m)\right)\delta_{n,m+1} this commutator is trace-class and hence compact, so that R𝒰R\in\mathcal{U}^{\mathcal{L}}. Moreover, clearly

indΛRk=k(k)\displaystyle\operatorname{ind}_{\Lambda}R^{k}=-k\qquad(k\in\mathbb{Z})

so that indΛ:𝒰\operatorname{ind}_{\Lambda}:\mathcal{U}^{\mathcal{L}}\to\mathbb{Z} is surjective.

We turn to surjectivity in the case 𝔽=\mathbb{F}=\mathbb{R}. We can use the same proof as above, being careful to make sure that the shift operator RR we define obeys [R,C]=0[R,C]=0. To do so, using the fact [Λ,C]=0[\Lambda,C]=0, we may let CC act on each subspace kerΛ,imΛ\ker\Lambda,\operatorname{im}\Lambda separately, and we invoke Lemma˜B.5 on each to obtain an ONB for each subspace, {φn}n0\Set{\varphi_{n}}_{n\leq 0} and {φn}n>0\Set{\varphi_{n}}_{n>0} respectively, which is fixed by CC. Defining again Rφi:=φi+1R\varphi_{i}:=\varphi_{i+1} then does the job.

We next turn to the case 𝔽=\mathbb{F}=\mathbb{H}. Let us begin with well-definededness of the index, i.e., we show that indΛ(U)2\operatorname{ind}_{\Lambda}(U)\in 2\mathbb{Z} for all U𝒰U\in\mathcal{U}_{\mathbb{H}}^{\mathcal{L}}. For such a UU, since [J,Λ]=0[J,\Lambda]=0, we may let JJ act separately on kerΛ,imΛ\ker\Lambda,\operatorname{im}\Lambda. Moreover, [U,J]=0[U,J]=0 implies [UL,J]=0[U_{L},J]=0 where ULΛUΛ:kerΛkerΛU_{L}\equiv\Lambda^{\perp}U\Lambda^{\perp}:\ker\Lambda\to\ker\Lambda. As such, we have JkerUL=kerULJ\ker U_{L}=\ker U_{L} and JkerUL=kerULJ\ker U_{L}^{\ast}=\ker U_{L}^{\ast}. As a result, these two spaces are even dimensional thanks to Lemma˜B.5 and hence the index is even.

Finally we turn to the surjectivity of indΛ:𝒰2\operatorname{ind}_{\Lambda}:\mathcal{U}^{\mathcal{L}}_{\mathbb{H}}\to 2\mathbb{Z}. Since [J,Λ]=0[J,\Lambda]=0, we may let JJ act separately on kerΛ,imΛ\ker\Lambda,\operatorname{im}\Lambda and invoke Lemma˜B.5 separately on them to obtain two orthonormal bases

(3.8) {φn,ψn}n0,{φn,ψn}n>0\displaystyle\Set{\varphi_{n},\psi_{n}}_{n\leq 0},\Set{\varphi_{n},\psi_{n}}_{n>0}

for kerΛ,imΛ\ker\Lambda,\operatorname{im}\Lambda respectively, which obey the property that Jφn=ψnJ\varphi_{n}=\psi_{n}. We then define a unitary operator RR on \mathcal{H} via

φnφn+1ψnψn+1(n)\displaystyle\varphi_{n}\mapsto\varphi_{n+1}\qquad\psi_{n}\mapsto\psi_{n+1}\qquad(n\in\mathbb{Z})

and note that

RJφi=Rψi=ψi+1=Jφi+1=JRφi\displaystyle RJ\varphi_{i}=R\psi_{i}=\psi_{i+1}=J\varphi_{i+1}=JR\varphi_{i}
RJψi=Rφi=φi+1=Jψi+1=JRψi.\displaystyle RJ\psi_{i}=-R\varphi_{i}=-\varphi_{i+1}=J\psi_{i+1}=JR\psi_{i}\,.

Thus RJ=JRRJ=JR so that R𝒰R\in\mathcal{U}^{\mathcal{L}}_{\mathbb{H}} and indΛRk=2k\operatorname{ind}_{\Lambda}R^{k}=-2k for all kk\in\mathbb{Z}.

We are left with establishing injectivity, which we can do for all three 𝔽=,,\mathbb{F}=\mathbb{C},\mathbb{R},\mathbb{H}. It is tantamount to the following statement: given any U,V𝒰𝔽U,V\in\mathcal{U}_{\mathbb{F}}^{\mathcal{L}} such that indΛU=indΛV\operatorname{ind}_{\Lambda}U=\operatorname{ind}_{\Lambda}V, there exists a (norm) continuous path

γ:[0,1]𝒰𝔽\displaystyle\gamma:[0,1]\to\mathcal{U}_{\mathbb{F}}^{\mathcal{L}}

such that γ(0)=U\gamma(0)=U and γ(1)=V\gamma(1)=V. Thanks to Lemma˜2.10 we may WLOG assume that V=𝟙V=\mathds{1} and hence that indΛU=0\operatorname{ind}_{\Lambda}U=0. But then an application of Lemma˜3.4 on UU yields

U=AB\displaystyle U=AB

for A,B𝒰𝔽A,B\in\mathcal{U}_{\mathbb{F}}^{\mathcal{L}} with 𝟙A𝒦\mathds{1}-A\in\mathcal{K} and [Λ,B]=0[\Lambda,B]=0. This first of all implies that AA can only have accumulation of spectrum at +1𝕊1+1\in\mathbb{S}^{1}. Now the analysis divides according to the value of 𝔽\mathbb{F}. In the simplest case, if 𝔽=\mathbb{F}=\mathbb{C}, let α𝕊1σ(A)\alpha\in\mathbb{S}^{1}\setminus\sigma(A). Then logα\log_{\alpha}, which is the holomorphic logarithm with branch cut at α\alpha, defines a local self-adjoint operator ilogα(U)-\operatorname{i}\log_{\alpha}(U). With that, Kuiper’s path eq.˜1.1 passes within 𝒰\mathcal{U}^{\mathcal{L}} thanks to Lemma˜2.8. For 𝔽=,\mathbb{F}=\mathbb{R},\mathbb{H}, the claim is shown via Lemma˜3.6 right below.

Next, since [B,Λ]=0[B,\Lambda]=0, we may write B=BLBRB=B_{L}\oplus B_{R} in the grading =(imΛ)imΛ\mathcal{H}=(\operatorname{im}\Lambda)^{\perp}\oplus\operatorname{im}\Lambda, for two unitaries BL,BR𝒰𝔽B_{L},B_{R}\in\mathcal{U}_{\mathbb{F}}. Now Theorem˜B.1 guarantees paths BL,BR𝟙B_{L},B_{R}\to\mathds{1} which pass within 𝒰𝔽\mathcal{U}_{\mathbb{F}}. Taking the direct sum of these two paths we obtain a diagonal, and hence local path B𝟙B\to\mathds{1} within 𝒰𝔽\mathcal{U}_{\mathbb{F}}^{\mathcal{L}}.

Combining the two separate paths from either A,BA,B to 𝟙\mathds{1} by multiplication yields the desired path. ∎

The following result which was used just above shows that when a local unitary has a gap, it may be deformed to the identity in a local way, i.e., Kuiper’s path eq.˜1.1 may be taken as local using Lemma˜2.8. Next, we establish this also in the presence of symmetries:

Lemma 3.6.

Let 𝔽{,}\mathbb{F}\in\Set{\mathbb{R},\mathbb{H}} and A𝒰𝔽A\in\mathcal{U}_{\mathbb{F}}^{\mathcal{L}} such that 𝟙A𝒦\mathds{1}-A\in\mathcal{K}. Then there exists a continuous path [0,1]tAt𝒰𝔽[0,1]\ni t\mapsto A_{t}\in\mathcal{U}_{\mathbb{F}}^{\mathcal{L}} such that A0=AA_{0}=A and A1=𝟙A_{1}=\mathds{1}.

Proof.

Since 𝟙A𝒦\mathds{1}-A\in\mathcal{K}, σ(A)𝕊1\sigma(A)\neq\mathbb{S}^{1}, and in fact, all the spectrum of AA outside of 11 consists of finitely degenerate eigenvalues.

Suppose 1σ(A)-1\notin\sigma(A), then there is a gap around 1-1 in σ(A)\sigma(A) and

At=pol((1t)A+t𝟙)\displaystyle A_{t}=\operatorname{pol}((1-t)A+t\mathds{1})

is the path we need. Indeed, the polar part AA|A|1A\mapsto A|A|^{-1} is a norm continuous mapping on operators that have a spectral gap about zero, which (1t)A+t𝟙(1-t)A+t\mathds{1} does for any t[0,1]t\in[0,1] thanks to 1σ(A)-1\notin\sigma(A). Moreover, pol\operatorname{pol} preserves symmetry by Lemma˜B.3 and locality by Lemma˜2.8. Indeed, |(1t)A+t𝟙|2|(1-t)A+t\mathds{1}|^{2} is clearly local, and λλ1/2\lambda\mapsto\lambda^{-1/2} is a continuous function.

Now assume 1σ(A)-1\in\sigma(A). Let V:=ker(A+𝟙)V:=\ker(A+\mathds{1}) denote the 1-1 eigenspace for AA. For brevity let F=C,JF=C,J according to the value of 𝔽\mathbb{F}. Since AF=FAAF=FA, if Aφ=φA\varphi=-\varphi, then AFφ=FAφ=FφAF\varphi=FA\varphi=-F\varphi, i.e., FφVF\varphi\in V iff φV\varphi\in V. Thus

FV=V,FV=V\displaystyle FV=V,\quad FV^{\perp}=V^{\perp}

Note the space VV is finite dimensional since 1-1 is in the discrete spectrum of AA. We decompose AA in =VV\mathcal{H}=V\oplus V^{\perp} as

A=[𝟙00AV]=[𝟙00𝟙][𝟙00AV]\displaystyle A=\begin{bmatrix}-\mathds{1}&0\\ 0&A_{V^{\perp}}\end{bmatrix}=\begin{bmatrix}-\mathds{1}&0\\ 0&\mathds{1}\end{bmatrix}\begin{bmatrix}\mathds{1}&0\\ 0&A_{V^{\perp}}\end{bmatrix}

Now A~:=𝟙AV\widetilde{A}:=\mathds{1}\oplus A_{V^{\perp}} belongs to 𝒰𝔽\mathcal{U}_{\mathbb{F}}^{\mathcal{L}} (as A𝒰𝔽A\in\mathcal{U}^{\mathcal{L}}_{\mathbb{F}} and (𝟙)𝟙(-\mathds{1})\oplus\mathds{1} is unitary, local, and commutes with FF) and 1σ(A~)-1\notin\sigma(\widetilde{A}), one can deform A~𝟙\widetilde{A}\to\mathds{1} within 𝒰𝔽\mathcal{U}_{\mathbb{F}}^{\mathcal{L}} as shown in the first paragraph.

We now deform (𝟙V)𝟙V(-\mathds{1}_{V})\oplus\mathds{1}_{V^{\perp}} to 𝟙-\mathds{1} within 𝒰𝔽\mathcal{U}_{\mathbb{F}}^{\mathcal{L}}. Since 𝟙V-\mathds{1}_{V} is already where it should be, we concentrate on deforming 𝟙V\mathds{1}_{V^{\perp}} to 𝟙V-\mathds{1}_{V^{\perp}} in a Λ\Lambda-local way. We decompose Λ\Lambda in =VV\mathcal{H}=V\oplus V^{\perp} as

Λ=[Λ11Λ12Λ21Λ22].\displaystyle\Lambda=\begin{bmatrix}\Lambda_{11}&\Lambda_{12}\\ \Lambda_{21}&\Lambda_{22}\end{bmatrix}\,.

Here Λ22\Lambda_{22} is essentially a projection (as VV is finite dimensional), and hence Lemma˜B.8 there exists a self-adjoint projection P:VVP:V^{\perp}\to V^{\perp} such that PΛ22𝒦P-\Lambda_{22}\in\mathcal{K}. In particular, [P,F]=0[P,F]=0 too (FF is diagonal in the VV decomposition). Now, to deform 𝟙V\mathds{1}_{V^{\perp}} to 𝟙V-\mathds{1}_{V^{\perp}} in a PP-local way, since 𝟙V\mathds{1}_{V^{\perp}} is diagonal in a PP-grading of VV^{\perp}, we may deform each diagonal block separately using Kuiper Theorem˜B.1 and that path is guaranteed to be PP-local as it is PP-diagonal. Let tWtt\mapsto W_{t} denote this deformation from 𝟙V\mathds{1}_{V^{\perp}} to 𝟙V-\mathds{1}_{V^{\perp}}. Then Wt𝒰𝔽W_{t}\in\mathcal{U}_{\mathbb{F}} and [Wt,P]=0[W_{t},P]=0. In particular

[Wt,Λ22]=[Wt,(Λ22P)+P]=[Wt,Λ22P]𝒦.\displaystyle[W_{t},\Lambda_{22}]=[W_{t},(\Lambda_{22}-P)+P]=[W_{t},\Lambda_{22}-P]\in\mathcal{K}\,.

Thus (𝟙)Wt𝒰𝔽(-\mathds{1})\oplus W_{t}\in\mathcal{U}_{\mathbb{F}}^{\mathcal{L}} deforms (𝟙)𝟙(-\mathds{1})\oplus\mathds{1} to 𝟙-\mathds{1} in a Λ\Lambda-local way. Now we can deform 𝟙-\mathds{1} to 𝟙\mathds{1} within 𝒰𝔽\mathcal{U}_{\mathbb{F}}^{\mathcal{L}} by decomposing 𝟙-\mathds{1} in the Λ\Lambda grading and the argument proceeds similarly. ∎

Next, we turn to the ,\star\mathbb{R},\star\mathbb{H} classes.

Proof of Theorem˜3.3.

We begin by establishing surjectivity (only for the case 𝔽=\mathbb{F}=\star\mathbb{H}, the other case being nullhomotopic). Clearly, we have 𝟙𝒰\mathds{1}\in\mathcal{U}_{\star\mathbb{H}}^{\mathcal{L}} and indΛ,2(𝟙)=0\operatorname{ind}_{\Lambda,2}(\mathds{1})=0. We are left to construct some U𝒰U\in\mathcal{U}_{\star\mathbb{H}}^{\mathcal{L}} with indΛ,2(U)=1\operatorname{ind}_{\Lambda,2}(U)=1. Using the same basis choice as in eq.˜3.8 we define a unitary operator UU on \mathcal{H} via

φnφn1,ψnψn+1(n)\displaystyle\varphi_{n}\mapsto\varphi_{n-1},\quad\psi_{n}\mapsto\psi_{n+1}\qquad(n\in\mathbb{Z})

and note that

UJφn=Uψn=ψn+1=Jφn+1=JUφn\displaystyle UJ\varphi_{n}=U\psi_{n}=\psi_{n+1}=J\varphi_{n+1}=JU^{*}\varphi_{n}
UJψn=Uφn=φn1=Jψn1=JUψn.\displaystyle UJ\psi_{n}=-U\varphi_{n}=-\varphi_{n-1}=J\psi_{n-1}=JU^{*}\psi_{n}\,.

Thus UJ=JUUJ=JU^{*} so that U𝒰U\in\mathcal{U}^{\mathcal{L}}_{\star\mathbb{H}}. In particular, since URRU\cong R\oplus R^{\ast} (RR being the right shift operator) in the decomposition span({φn}n)span({ψn}n)\operatorname{span}(\Set{\varphi_{n}}_{n})\oplus\operatorname{span}(\Set{\psi_{n}}_{n}), dimkerΛRR=1\dim\ker\mathbb{\Lambda}R\oplus R^{\ast}=1 and hence indΛ,2(U)=1\operatorname{ind}_{\Lambda,2}(U)=1.

Next, we deal with the proof of injectivity for both 𝔽=,\mathbb{F}=\star\mathbb{R},\star\mathbb{H} (with F=C,JF=C,J accordingly). Let U𝒰𝔽U\in\mathcal{U}_{\mathbb{F}}^{\mathcal{L}} and if 𝔽=\mathbb{F}=\star\mathbb{H}, assume for the moment that indΛ,2U=0\operatorname{ind}_{\Lambda,2}U=0, the other case will be dealt with separately. Applying Lemma˜3.5 on UU yields

U=AB\displaystyle U=AB

where A𝒰A\in\mathcal{U}^{\mathcal{L}} with A(BF)=(BF)AA(BF)=(BF)A^{*} and 𝟙A𝒦\mathds{1}-A\in\mathcal{K}, and B𝒰𝔽B\in\mathcal{U}_{\mathbb{F}}^{\mathcal{L}} and [Λ,B]=0[\Lambda,B]=0. Since 𝟙A𝒦\mathds{1}-A\in\mathcal{K}, then the spectrum of AA can only accumulate at +1+1. We may rotate, e.g., anti-clockwise by θ\theta degree, the spectrum σ(A)𝕊1\sigma(A)\subset\mathbb{S}^{1} of AA so that there is a gap at 1-1. In particular eiθA𝒰𝔽e^{i\theta}A\in\mathcal{U}_{\mathbb{F}} since FF is anti-\mathbb{C}-linear, so

eiθAF=FeiθA=F(eiθA)\displaystyle e^{i\theta}AF=Fe^{-i\theta}A^{\ast}=F(e^{i\theta}A)^{*}

Thus WLOG we may assume there is a gap at 1-1 in the spectrum of AA. Now consider the path

[0,1]tAt=pol((1t)A+t𝟙).\displaystyle[0,1]\ni t\mapsto A_{t}=\operatorname{pol}((1-t)A+t\mathds{1})\,.

Then At𝒰𝔽A_{t}\in\mathcal{U}^{\mathcal{L}}_{\mathbb{F}} with respect to BFBF by Lemma˜B.3. Consider tAtBt\mapsto A_{t}B which deforms UU to BB. We have

AtBF=BFAt=FBAt=F(AtB).\displaystyle A_{t}BF=BFA_{t}^{*}=FB^{*}A_{t}^{*}=F(A_{t}B)^{*}\,.

Thus AtB𝒰𝔽A_{t}B\in\mathcal{U}_{\mathbb{F}}^{\mathcal{L}} for all tt. Finally, since [B,Λ]=0[B,\Lambda]=0, we may use Theorem˜B.2 to deform each diagonal block of BB in the Λ\Lambda grading to 𝟙\mathds{1}, resulting in B𝟙B\to\mathds{1} within 𝒰𝔽\mathcal{U}_{\mathbb{F}}^{\mathcal{L}}.

Finally, consider 𝔽=\mathbb{F}=\star\mathbb{H} assuming that indΛ,2U=1\operatorname{ind}_{\Lambda,2}U=1. Unlike in the ordinary \mathbb{Z}-valued index where we use the logarithmic law Lemma˜2.10 and then we may always connect a zero-index operator to 𝟙\mathds{1}, in the present case for the 2\mathbb{Z}_{2} index, we rather directly argue by connecting any two non-zero index operators together. Hence consider U,U~𝒰U,\widetilde{U}\in\mathcal{U}_{\star\mathbb{H}}^{\mathcal{L}} both with non-zero index.

Let us start by deforming UU (and also U~\widetilde{U}) to a more convenient non-zero index operator. Decompose UU in block form in the Λ\Lambda grading as before in eq.˜3.6. For Z=ULLZ=U_{LL} for brevity and X=pol(Z)X=\operatorname{pol}(Z)\in\mathcal{B}_{\star\mathbb{H}}, let mm be defined by 2m+1:=dimkerX=dim(imX)2m+1:=\dim\ker X=\dim(\operatorname{im}X)^{\perp}. We use ˜A.4 to conclude that dimkerURR\dim\ker U_{RR} is odd as well. Similar to the proof in Lemma˜3.5, we may extend XX to a partial isometry YY\in\mathcal{B}_{\star\mathbb{H}} that has dimkerY=dim(imY)=1\dim\ker Y=\dim(\operatorname{im}Y)^{\perp}=1, and such that YZ𝒦Y-Z\in\mathcal{K}. Indeed, we cannot extend YY to a unitary in 𝒰\mathcal{U}_{\star\mathbb{H}} precisely since the kernels are not even dimensional (see Lemma˜B.6 whose hypothesis is a zero index). Let BL:=YB_{L}:=Y and BRB_{R} the corresponding construction out of URRU_{RR}. Even though we cannot extend BiB_{i} to a unitary separately, we will show that this may be done on the bigger space for the direct sum B:=BLBRB:=B_{L}\oplus B_{R}. Let kerBi\ker B_{i} be spanned by the unit vector ηi\eta_{i}. Since BiB_{i}\in\mathcal{B}_{\star\mathbb{H}}, (imBi)(\operatorname{im}B_{i})^{\perp} is spanned by the vector ξi:=Jηi\xi_{i}:=J\eta_{i}. We now define an operator MM, which is unitary when considered as a map M:kerB(imB)M:\ker B\to(\operatorname{im}B)^{\perp}, written in the Λ\Lambda-grading as

M:=[0ξLηRξRηL0]\displaystyle M:=\begin{bmatrix}0&\xi_{L}\otimes\eta_{R}^{*}\\ -\xi_{R}\otimes\eta_{L}^{*}&0\end{bmatrix}

that satisfies MJ=JMMJ=JM^{*}. Indeed, we have

MJ\displaystyle MJ =[0ξLηRξRηL0][J00J]\displaystyle=\begin{bmatrix}0&\xi_{L}\otimes\eta_{R}^{*}\\ -\xi_{R}\otimes\eta_{L}^{*}&0\end{bmatrix}\begin{bmatrix}J&0\\ 0&J\end{bmatrix}
=[0(ξLηR)J(ξRηL)J0]\displaystyle=\begin{bmatrix}0&(\xi_{L}\otimes\eta_{R}^{*})J\\ (-\xi_{R}\otimes\eta_{L}^{*})J&0\end{bmatrix}
=[0J(ηLξR)J(ηRξL)0]\displaystyle=\begin{bmatrix}0&J(-\eta_{L}\otimes\xi_{R}^{*})\\ J(\eta_{R}\otimes\xi_{L}^{*})&0\end{bmatrix}
=JM\displaystyle=JM^{*}

where in the third equality, we use the fact

(ξLηR)Jφ=(ηR,Jφ)ξL=(φ,JηR)ξL=(φ,ξR)ξL=J(ξR,φ)ηL=J(ηLξR)φ\displaystyle(\xi_{L}\otimes\eta_{R}^{*})J\varphi=(\eta_{R},J\varphi)\xi_{L}=(\varphi,J^{*}\eta_{R})\xi_{L}=-(\varphi,\xi_{R})\xi_{L}=-J(\xi_{R},\varphi)\eta_{L}=J(-\eta_{L}\otimes\xi_{R}^{*})\varphi

and similarly for (ξRηL)J=J(ηRξL).(-\xi_{R}\otimes\eta_{L}^{*})J=J(\eta_{R}\otimes\xi_{L}^{*}). With MM we define the operator

(3.9) D:=B+M=[BLξLηRξRηLBR].\displaystyle D:=B+M=\begin{bmatrix}B_{L}&\xi_{L}\otimes\eta_{R}^{*}\\ -\xi_{R}\otimes\eta_{L}^{*}&B_{R}\end{bmatrix}\,.

Thus DD is unitary by construction, and local, since ξLηR\xi_{L}\otimes\eta_{R}^{*} and ξRηL-\xi_{R}\otimes\eta_{L}^{*} are finite-rank, and obeys DJ=JDDJ=JD^{\ast}. We define A:=UDA:=UD^{*}, write U=ADU=AD, and deform AA away using similar argument as before (𝟙A=(DU)D𝒦\mathds{1}-A=(D-U)D^{\ast}\in\mathcal{K} and A(DJ)=(DJ)AA(DJ)=(DJ)A^{\ast}). Thus there is a path from UU to DD within 𝒰\mathcal{U}^{\mathcal{L}}_{\star\mathbb{H}}, and similarly a path from U~\widetilde{U} to D~\widetilde{D}, where D~\widetilde{D} is constructed analogously to D.D.

Hence we are left to deform DD to D~\widetilde{D}. For each i{L,R}i\in\Set{L,R}, we use Lemma˜3.7 right below to construct unitaries Xi,YiX_{i},Y_{i} such that Bi=XiB~iYiB_{i}=X_{i}^{*}\widetilde{B}_{i}Y_{i} and YiJ=JXiY_{i}J=JX_{i}. Plugging in X,YX,Y into DD we find

D\displaystyle D =[XLB~LYLξLηRξRηLXRB~RYR]\displaystyle=\begin{bmatrix}X^{*}_{L}\widetilde{B}_{L}Y_{L}&\xi_{L}\otimes\eta_{R}^{*}\\ -\xi_{R}\otimes\eta_{L}^{*}&X_{R}^{*}\widetilde{B}_{R}Y_{R}\end{bmatrix}
=[XL00XR][B~LXL(ξLηR)YRXR(ξRηL)YLB~R][YL00YR].\displaystyle=\begin{bmatrix}X_{L}^{*}&0\\ 0&X_{R}^{*}\end{bmatrix}\begin{bmatrix}\widetilde{B}_{L}&X_{L}(\xi_{L}\otimes\eta_{R}^{*})Y_{R}^{*}\\ X_{R}(-\xi_{R}\otimes\eta_{L}^{*})Y_{L}^{*}&\widetilde{B}_{R}\end{bmatrix}\begin{bmatrix}Y_{L}&0\\ 0&Y_{R}\end{bmatrix}\,.

In particular, from how Xi,YiX_{i},Y_{i} are constructed in eq.˜3.12, one has

XR(ξRηL)YL=(XRξR)(YLηL)=ξ~Rη~L.\displaystyle X_{R}(-\xi_{R}\otimes\eta_{L}^{*})Y_{L}^{*}=-(X_{R}\xi_{R})\otimes(Y_{L}\eta_{L})^{*}=-\widetilde{\xi}_{R}\otimes\widetilde{\eta}_{L}^{*}\,.

Similarly XL(ξLηR)YR=ξ~Lη~RX_{L}(\xi_{L}\otimes\eta_{R}^{*})Y_{R}^{*}=\tilde{\xi}_{L}\otimes\widetilde{\eta}_{R}^{*}. We write X=XLXRX=X_{L}\oplus X_{R} and Y=YLYRY=Y_{L}\oplus Y_{R}. Then

D=XD~Y=XD~JXJ\displaystyle D=X^{*}\widetilde{D}Y=X^{*}\widetilde{D}JXJ^{*}

where we used YJ=JXYJ=JX in the last step. Applying Theorem˜B.1 on XX, there exists a path [0,1]tXt𝒰[0,1]\ni t\mapsto X_{t}\in\mathcal{U} connecting X𝟙X\rightsquigarrow\mathds{1}. Let Bt=XtB~JXtJB_{t}=X_{t}^{*}\widetilde{B}JX_{t}J^{*}. Then

BtJ=XtB~JXt=XtJB~Xt=J(JXtJB~Xt)=JBt.\displaystyle B_{t}J=X_{t}^{*}\widetilde{B}JX_{t}=X_{t}^{*}J\widetilde{B}^{*}X_{t}=J(JX_{t}^{*}J^{*}\widetilde{B}^{*}X_{t})=JB_{t}^{*}\,.

Thus Bt𝒰B_{t}\in\mathcal{U}_{\star\mathbb{H}}^{\mathcal{L}} deforms BB to B~\widetilde{B} as desired.

Lemma 3.7.

Let U,U~U,\widetilde{U}\in\mathcal{B}_{\star\mathbb{H}} be partial isometries such that the dimension of the kernels of U,U~,U,U~U,\widetilde{U},U^{*},\widetilde{U}^{*} are all finite and equal. Then there exists V,W𝒰V,W\in\mathcal{U} such that U=VU~WU=V^{*}\widetilde{U}W and WJ=JV.WJ=JV.

Proof.

We first write

U\displaystyle U =[000U2]:kerU(kerU)(imU)imU\displaystyle=\begin{bmatrix}0&0\\ 0&U_{2}\end{bmatrix}:\ker U\oplus(\ker U)^{\perp}\to(\operatorname{im}U)^{\perp}\oplus\operatorname{im}U
U~\displaystyle\widetilde{U} =[000U~2]:kerU~(kerU~)(imU~)imU~.\displaystyle=\begin{bmatrix}0&0\\ 0&\widetilde{U}_{2}\end{bmatrix}:\ker\widetilde{U}\oplus(\ker\widetilde{U})^{\perp}\to(\operatorname{im}\widetilde{U})^{\perp}\oplus\operatorname{im}\widetilde{U}\,.

Define operators V,WV,W that take the block form

W\displaystyle W =[W100W2]:kerU(kerU)kerU~(kerU~)\displaystyle=\begin{bmatrix}W_{1}&0\\ 0&W_{2}\end{bmatrix}:\ker U\oplus(\ker U)^{\perp}\to\ker\widetilde{U}\oplus(\ker\widetilde{U})^{\perp}
V\displaystyle V =[V100U~2W2U2]:(imU)imU(imU~)imU~\displaystyle=\begin{bmatrix}V_{1}&0\\ 0&\widetilde{U}_{2}W_{2}U_{2}^{*}\end{bmatrix}:(\operatorname{im}U)^{\perp}\oplus\operatorname{im}U\to(\operatorname{im}\widetilde{U})^{\perp}\oplus\operatorname{im}\widetilde{U}

where V1,W1,W2V_{1},W_{1},W_{2} are, for now, some unitaries to be constructed explicitly momentarily. With this, it is clear that

U=VU~W.\displaystyle U=V^{*}\widetilde{U}W\,.

We construct the unitaries V1,W1,W2V_{1},W_{1},W_{2} such that the following two conditions hold

(3.10) W1J\displaystyle W_{1}J =JV1\displaystyle=JV_{1}
(3.11) W2J\displaystyle W_{2}J =JU~2W2U2.\displaystyle=J\widetilde{U}_{2}W_{2}U_{2}^{*}\,.

The expressions make sense since UU\in\mathcal{B}_{\star\mathbb{H}} implies that

JkerU=(imU),J(kerU)=imU;\displaystyle J\ker U=(\operatorname{im}U)^{\perp}\,,\quad J(\ker U)^{\perp}=\operatorname{im}U\,;

similar expressions holds for U~\widetilde{U}. Now we let kerU\ker U be spanned by the ONB {ηi}i=1m\Set{\eta_{i}}_{i=1}^{m}. Then (imU)(\operatorname{im}U)^{\perp} is spanned by the ONB {ξi}i=1m\Set{\xi_{i}}_{i=1}^{m}, where ξi:=Jηi\xi_{i}:=J\eta_{i}. We construct analogous tilde version of ONB for kerU~\ker\widetilde{U} and (imU~)(\operatorname{im}\widetilde{U})^{\perp}. Define

(3.12) V1:i=1mξ~ξi,W1:i=1mη~iηi.\displaystyle V_{1}:\sum_{i=1}^{m}\widetilde{\xi}\otimes\xi_{i}^{*},\quad W_{1}:\sum_{i=1}^{m}\widetilde{\eta}_{i}\otimes\eta_{i}^{*}.

Thus eq.˜3.10 holds.

We find that JU2JU_{2} defines a \star\mathbb{H}-structure on (kerU)(\ker U)^{\perp}, so applying Lemma˜B.5 gives an ONB consisting of Kramers pairs {φi,ψi}i=1\Set{\varphi_{i},\psi_{i}}_{i=1}^{\infty} for (kerU)(\ker U)^{\perp} such that ψi=JU2φi\psi_{i}=JU_{2}\varphi_{i}. Similarly, let {φi,ψ~i}i=1\Set{\varphi_{i},\widetilde{\psi}_{i}}_{i=1}^{\infty} be an ONB of Kramers pairs for JU~2J\widetilde{U}_{2} on (kerU~)(\ker\widetilde{U})^{\perp}. Construct W2W_{2} as

(3.13) φiψ~i,ψiφ~i\displaystyle\varphi_{i}\mapsto-\widetilde{\psi}_{i},\quad\psi_{i}\mapsto\widetilde{\varphi}_{i}

One again readily verifies that eq.˜3.11 holds. The relations eqs.˜3.10 and 3.11 are equivalent to

WJ=JV.\displaystyle WJ=JV\,.

4 Equivariant classification of Λ\Lambda-non-trivial self-adjoint unitaries

In this section we turn our attention to equivariant local self-adjoint (orthogonal) projections, and calculate the corresponding set of path-connected components. Now, however, we add a non-triviality condition which is stronger than locality, and moreover, the symmetry classes 𝔽\mathbb{F} we consider are slightly different. To explain the difference, let us consider equivariant local self-adjoint unitaries 𝒮𝔽\mathcal{S}^{\mathcal{L}}_{\mathbb{F}} rather than projections (see ˜2.4); we shall abbreviate SAU henceforth. We prefer working with SAUs here for two reasons: (1) the physical symmetry constraints appear naturally at the level of the self-adjoint unitaries, which, as we’ll see below, are flat Hamiltonians, and (2) the calculations below somewhat simplify in this way. As for the symmetries, we have still \mathbb{C} (no constraint), \mathbb{R} and \mathbb{H}, i.e., the SAU would commute with F=C,JF=C,J. However now we replace ,\star\mathbb{R},\star\mathbb{H} by i,i\operatorname{i}\mathbb{R},\operatorname{i}\mathbb{H} respectively (since these are the conditions which arise from particle-hole symmetry later, see Section˜5).

As was mentioned, now we constrain the class of SAUs we study even more beyond locality in a crucial way. We have already seen the notion of a non-trivial SAU in ˜2.1: this is a SAU operator where both ±1\pm 1 eigenspaces are infinite dimensional. We shall also need

Definition 4.1 (Λ\Lambda-non-trivial SAUs).

U𝒮U\in\mathcal{S} is called Λ\Lambda-non-trivial iff there exists some V𝒮V\in\mathcal{S} such that:

  1. 1.

    [V,Λ]=0[V,\Lambda]=0 (hyper-local).

  2. 2.

    ΛVΛ\Lambda V\Lambda and ΛVΛ\Lambda^{\perp}V\Lambda^{\perp} are both non-trivial SAUs.

  3. 3.

    UV𝒦U-V\in\mathcal{K}.

We note this implies automatically that such a UU is Λ\Lambda-local since by definition ΛUΛ\Lambda U\Lambda^{\perp} is compact. We denote the space of all Λ\Lambda-non-trivial SAUs by 𝒮Λnt\mathcal{S}^{\Lambda\mathrm{nt}}. Hence we have

𝒮Λnt𝒮𝒮.\displaystyle\mathcal{S}^{\Lambda\mathrm{nt}}\subsetneq\mathcal{S}^{\mathcal{L}}\subsetneq\mathcal{S}\,.

It turns out that if one attempts to classify the bare, merely Λ\Lambda-local space 𝒮\mathcal{S}^{\mathcal{L}}, the result is not nullhomotopic (as one would expect from Table˜1) due to finite rank problems which roughly correspond to half-infinite systems. So later on, in Section˜5 we will see that the correct notion to reproduce Table˜1 is rather the more constrained Λ\Lambda-non-trivial space 𝒮Λnt\mathcal{S}^{\Lambda\mathrm{nt}} and that in a sense, unitaries are automatically Λ\Lambda-non-trivial (see Lemma˜5.14 below), which is why this notion was not necessary in Section˜3. Another point in support of this notion is that Λ\Lambda-non-triviality is well-defined in the sense that it is preserved under small norm and compact perturbations within 𝒮\mathcal{S}^{\mathcal{L}}, see Lemma˜4.4 at the end of this section.

π0(𝒮Λnt)\pi_{0}(\mathcal{S}^{\Lambda\mathrm{nt}}) {0}\Set{0}
π0(𝒮Λnt)\pi_{0}(\mathcal{S}^{\Lambda\mathrm{nt}}_{\mathbb{R}}) {0}\Set{0}
π0(𝒮Λnt)\pi_{0}(\mathcal{S}^{\Lambda\mathrm{nt}}_{\mathbb{H}}) {0}\Set{0}
π0(𝒮iΛnt)\pi_{0}(\mathcal{S}^{\Lambda\mathrm{nt}}_{\operatorname{i}\mathbb{R}}) 2\mathbb{Z}_{2}
π0(𝒮iΛnt)\pi_{0}(\mathcal{S}^{\Lambda\mathrm{nt}}_{\operatorname{i}\mathbb{R}}) {0}\Set{0}
Table 3: The classification of equivariant non-trivial self-adjoint unitaries.

We finally turn to our main classification theorems. The results of this section are summarized in Table˜3.

Theorem 4.2 (Classification of Λ\Lambda-non-trivial ,,\mathbb{R},\mathbb{C},\mathbb{H} SAUs).

The space of Λ\Lambda-non-trivial ,,\mathbb{R},\mathbb{C},\mathbb{H} SAUs is null-homotopic:

(4.1) π0(𝒮𝔽Λnt){0}(𝔽{,,}).\displaystyle\pi_{0}(\mathcal{S}^{\Lambda\mathrm{nt}}_{\mathbb{F}})\cong\Set{0}\qquad(\mathbb{F}\in\Set{\mathbb{C},\mathbb{R},\mathbb{H}})\,.

When 𝔽=\mathbb{F}=\mathbb{C} (i.e., without any symmetry constraints) this theorem is not new, and appeared relatively recently within [ACL15]. Here we extend it also to the cases 𝔽=,\mathbb{F}=\mathbb{R},\mathbb{H} (that extension is straightforward) in a unified proof for all three cases.

The following theorem is new to our knowledge:

Theorem 4.3 (Classification of Λ\Lambda-non-trivial i\operatorname{i}\mathbb{R} and i\operatorname{i}\mathbb{H} SAUs).

The space of Λ\Lambda-non-trivial i\operatorname{i}\mathbb{R} SAUs has two path components. The map indΛ,2:𝒮iΛnt2\operatorname{ind}_{\Lambda,2}:\mathcal{S}^{\Lambda\mathrm{nt}}_{\operatorname{i}\mathbb{R}}\to\mathbb{Z}_{2} is norm continuous and ascends to a bijection

(4.2) indΛ,2:π0(𝒮iΛnt)2.\displaystyle\operatorname{ind}_{\Lambda,2}:\pi_{0}(\mathcal{S}^{\Lambda\mathrm{nt}}_{\operatorname{i}\mathbb{R}})\xrightarrow{\sim}\mathbb{Z}_{2}\,.

The space of Λ\Lambda-non-trivial i\operatorname{i}\mathbb{H} SAUs is null-homotopic:

(4.3) π0(𝒮iΛnt){0}.\displaystyle\pi_{0}(\mathcal{S}^{\Lambda\mathrm{nt}}_{\operatorname{i}\mathbb{H}})\cong\Set{0}\,.

We now present the proofs of Theorems˜4.2 and 4.3.

Proof of Theorem˜4.2.

Let U,U~𝒮𝔽ΛntU,\widetilde{U}\in\mathcal{S}^{\Lambda\mathrm{nt}}_{\mathbb{F}} be given, with 𝔽=,,\mathbb{F}=\mathbb{C},\mathbb{R},\mathbb{H}. Our goal is to construct a continuous path within 𝒮𝔽Λnt\mathcal{S}^{\Lambda\mathrm{nt}}_{\mathbb{F}} from UU to U~\widetilde{U}. The main idea is as follows: using Lemma˜B.7 below we know that non-trivial SAUs are null-homotopic without the further Λ\Lambda-non-triviality constraint. So if it turned out that both U,U~U,\widetilde{U} were diagonal in the Λ\Lambda grading, we would be finished. We thus concentrate on showing how UU may be deformed into a Λ\Lambda-diagonal element in 𝒮𝔽Λnt\mathcal{S}^{\Lambda\mathrm{nt}}_{\mathbb{F}}, with the deformation within that space.

Hence, in the =im(Λ)im(Λ)\mathcal{H}=\operatorname{im}(\Lambda)^{\perp}\oplus\operatorname{im}(\Lambda) grading, let us write

(4.4) U=[XAAY]\displaystyle U=\begin{bmatrix}X&A\\ A^{\ast}&Y\end{bmatrix}

where X,YX,Y are self-adjoint operators and A:im(Λ)im(Λ)A:\operatorname{im}(\Lambda)\to\operatorname{im}(\Lambda)^{\perp} is general. It should be emphasized that X,YX,Y are not SAUs since A0A\neq 0. By locality however, AA is compact, and since for any SAU, 𝟙U𝟙-\mathds{1}\leq U\leq\mathds{1}, we also have 𝟙X,Y𝟙-\mathds{1}\leq X,Y\leq\mathds{1}. Since 𝟙=U2\mathds{1}=U^{2}, we have X,YX,Y essentially unitary; in particular, |A|2=𝟙Y2|A|^{2}=\mathds{1}-Y^{2}, |A|2=𝟙X2|A^{\ast}|^{2}=\mathds{1}-X^{2}. Finally, the intertwining property

(4.5) XA=AY\displaystyle XA=-AY

holds. Using it, we can show that

(4.6) A:ker(Yλ𝟙)ker(X+λ𝟙)(λ(1,1))\displaystyle A:\ker(Y-\lambda\mathds{1})\to\ker(X+\lambda\mathds{1})\qquad(\lambda\in(-1,1))

isomorphically. Indeed, let Yψ=λψY\psi=\lambda\psi for |λ|<1|\lambda|<1 and ψ=1\left\lVert\psi\right\rVert=1. Then XAψ=λAψXA\psi=-\lambda A\psi and

(4.7) Aψ2=ψ,|A|2ψ=ψ,(𝟙Y2)ψ=(1λ2)ψ20.\displaystyle\left\lVert A\psi\right\rVert^{2}=\langle\psi,|A|^{2}\psi\rangle=\langle\psi,(\mathds{1}-Y^{2})\psi\rangle=(1-\lambda^{2})\left\lVert\psi\right\rVert^{2}\neq 0\,.

This works similarly to show that AA^{\ast} is injective and hence AA is the claimed isomorphism.

Now, X,YX,Y have spectra which may only accumulate at ±1\pm 1, since

σess(U)=σess(X)σess(Y)\displaystyle\sigma_{\mathrm{ess}}(U)=\sigma_{\mathrm{ess}}(X)\cup\sigma_{\mathrm{ess}}(Y)

so that on (1,1)(-1,1) both XX and YY have discrete spectrum, and the intertwining property implies that

(4.8) χ{λ}(X)A=Aχ{λ}(Y)\displaystyle\chi_{\Set{\lambda}}(X)A=A\chi_{\Set{-\lambda}}(Y)

for all λ(1,1)\lambda\in(-1,1). Indeed,

Aχ{λ}(Y)=χ{λ}(X)Aχ{λ}(Y)=χ{λ}(X)A(𝟙χ{λ}(Y))\displaystyle A\chi_{\Set{-\lambda}}(Y)=\chi_{\Set{\lambda}}(X)A\chi_{\Set{-\lambda}}(Y)=\chi_{\Set{\lambda}}(X)A\left(\mathds{1}-\chi_{\Set{-\lambda}}(Y)^{\perp}\right)

where the first inequality is thanks to eq.˜4.6. Now, χ{λ}(X)Aχ{λ}(Y)=0\chi_{\Set{\lambda}}(X)A\chi_{\Set{-\lambda}}(Y)^{\perp}=0 because ker(A)=ker(|A|2)=ker(𝟙Y2)\ker(A)=\ker(|A|^{2})=\ker(\mathds{1}-Y^{2}), so AA is zero on any eigenvector of YY of eigenvalue of modulus 11, and on any other eigenvalue μ-\mu, eq.˜4.6 shows that it maps to the range of χ{μ}(X)\chi_{\Set{-\mu}}(X).

We will use the definition of sgn:\operatorname{sgn}:\mathbb{R}\to\mathbb{R} given by

(4.9) sgn(λ)χ(,0)(λ)+χ(0,)(λ)(λ)\displaystyle\operatorname{sgn}(\lambda)\equiv-\chi_{(-\infty,0)}(\lambda)+\chi_{(0,\infty)}(\lambda)\qquad(\lambda\in\mathbb{R})

and with it define f±:f_{\pm}:\mathbb{R}\to\mathbb{R} via

(4.10) f±(λ):=sgn(λ)±χ{0}(λ)(λ).\displaystyle f_{\pm}(\lambda):=\operatorname{sgn}(\lambda)\pm\chi_{\Set{0}}(\lambda)\qquad(\lambda\in\mathbb{R})\,.

We define the diagonal SAU operator

(4.11) V:=f+(X)f(Y).\displaystyle V:=f_{+}(X)\oplus f_{-}(Y)\,.

Here V:=f+(X)f+(Y)V:=f_{+}(X)\oplus f_{+}(Y) also works. This is the operator we shall deform UU into. To do so, we shall use the conjugating self-adjoint operator due to [ACL15]:

(4.12) G:=12(U+V).\displaystyle G:=\frac{1}{2}\left(U+V\right)\,.

Let us note the effect of different symmetries in this context, i.e., we assume that [F,U]=0[F,U]=0 for F=𝟙,C,JF=\mathds{1},C,J. Since the symmetry operators are hyper-local by Assumption˜2.11, we have X,Y𝔽X,Y\in\mathcal{B}_{\mathbb{F}} too. In particular sgn(X),sgn(Y)\operatorname{sgn}(X),\operatorname{sgn}(Y) and χ{0}(X),χ{0}(Y)\chi_{\Set{0}}(X),\chi_{\Set{0}}(Y) also belong to 𝔽.\mathcal{B}_{\mathbb{F}}. Indeed, even though FF is anti-unitary, X,YX,Y are self-adjoint and hence the anti-unitarity does not interfere. Hence G𝔽G\in\mathcal{B}_{\mathbb{F}} too, and so we can essentially forget about the symmetry constraints as long as we have symmetric versions of Theorems˜B.1 and B.7, which we do.

Next, we note that that since U2=V2=𝟙U^{2}=V^{2}=\mathds{1}, GU=12(U+V)U=12V(U+V)=VGGU=\frac{1}{2}(U+V)U=\frac{1}{2}V(U+V)=VG. Also, GG is clearly Fredholm of zero Λ\Lambda-index (in the sense of eq.˜2.4). Indeed, since λλ+f±(λ)=:g±(λ)\lambda\mapsto\lambda+f_{\pm}(\lambda)=:g_{\pm}(\lambda) has im(g±)(1,1)=\operatorname{im}(g_{\pm})\cap(-1,1)=\varnothing, GG is, up to compact, a direct sum of invertible, and hence zero-index Fredholm, operators:

G12(g+(X)g(Y))𝒦.\displaystyle G-\frac{1}{2}\left(g_{+}(X)\oplus g_{-}(Y)\right)\in\mathcal{K}\,.

We find

indΛG=indg+(X)=0.\displaystyle\operatorname{ind}_{\Lambda}G=\operatorname{ind}g_{+}(X)=0\,.

In fact G=12[g+(X)AAg(Y)]G=\frac{1}{2}\begin{bmatrix}g_{+}(X)&A\\ A^{\ast}&g_{-}(Y)\end{bmatrix} is invertible. Indeed, since indG=0\operatorname{ind}G=0, it suffices to check that kerG={0}{\ker G=\Set{0}}. Suppose G[φψ]=0G\begin{bmatrix}\varphi\\ \psi\end{bmatrix}=0, then

(4.13) g+(X)φ+Aψ=0\displaystyle g_{+}(X)\varphi+A\psi=0
(4.14) Aφ+g(Y)ψ=0.\displaystyle A^{*}\varphi+g_{-}(Y)\psi=0\,.

Multiply the first equation by AA^{*} to get

Ag+(X)φ+|A|2ψ=0.\displaystyle A^{*}g_{+}(X)\varphi+|A|^{2}\psi=0\,.

Now, using the intertwining property eq.˜4.5, we have

Ag+(X)\displaystyle A^{*}g_{+}(X) =A(X+sgn(X)+χ{0}(X))\displaystyle=A^{*}(X+\operatorname{sgn}(X)+\chi_{\Set{0}}(X))
=(Ysgn(Y)+χ{0}(Y))A=g(Y)A.\displaystyle=(-Y-\operatorname{sgn}(Y)+\chi_{\Set{0}}(Y))A^{*}=-g_{-}(Y)A^{*}\,.

Thus

0=Ag+(X)φ+|A|2ψ=g(Y)Aφ+|A|2ψ=(g2(Y)+𝟙Y2)ψ\displaystyle 0=A^{*}g_{+}(X)\varphi+|A|^{2}\psi=-g_{-}(Y)A^{*}\varphi+|A|^{2}\psi=(g_{-}^{2}(Y)+\mathds{1}-Y^{2})\psi

where, in the last equality, we have used eq.˜4.14, and |A|2=𝟙Y2.|A|^{2}=\mathds{1}-Y^{2}. Note that λg2(λ)+1λ2=2|λ|+2\lambda\mapsto g_{-}^{2}(\lambda)+1-\lambda^{2}=2|\lambda|+2 has range in [2,)[2,\infty), and hence the above implies that ψ=0.\psi=0. Similarly, one can show that φ=0.\varphi=0.

We now use the self-adjoint invertible operator GG as follows. Since GU=VGGU=VG, we have G2U=UG2G^{2}U=UG^{2} and thus also |G|U=U|G||G|U=U|G|. This implies that we also have pol(G)U=Vpol(G)\operatorname{pol}(G)U=V\operatorname{pol}(G), i.e.,

(4.15) V=pol(G)Upol(G).\displaystyle V=\operatorname{pol}(G)^{\ast}U\operatorname{pol}(G)\,.

But pol(G)𝒰𝔽\operatorname{pol}(G)\in\mathcal{U}^{\mathcal{L}}_{\mathbb{F}} (we are using Lemma˜B.3) and, indΛpol(G)=0\operatorname{ind}_{\Lambda}\operatorname{pol}(G)=0 (the index is invariant under the taking the polar part [GS18, Lemma 6]). Thus using Theorem˜3.1 pol(G)\operatorname{pol}(G) may be deformed to 𝟙\mathds{1} within the space 𝒰𝔽\mathcal{U}^{\mathcal{L}}_{\mathbb{F}}. This yields an equivariant local path within 𝒰𝔽\mathcal{U}_{\mathbb{F}}

UVLemma B.7V~U~.\displaystyle U\rightsquigarrow V\overset{\mathrm{\lx@cref{creftype~refnum}{lem:classification of nontrivial projections}}}{\rightsquigarrow}\widetilde{V}\rightsquigarrow\widetilde{U}\,.

So far we have only exhibited this path as Λ\Lambda-local, but it is in fact Λ\Lambda-non-trivial via Lemma˜4.4 below.

We turn to the two remaining, more exotic symmetry classes, i\operatorname{i}\mathbb{R} and i\operatorname{i}\mathbb{H}, one of which is not nullhomotopic.

Proof of Theorem˜4.3.

Since we claim that 𝒮iΛnt\mathcal{S}^{\Lambda\mathrm{nt}}_{\operatorname{i}\mathbb{H}} is nullhomotopic, we only need to establish surjectivity for ind2,Λ:𝒮iΛnt2\operatorname{ind}_{2,\Lambda}:\mathcal{S}^{\Lambda\mathrm{nt}}_{\operatorname{i}\mathbb{R}}\to\mathbb{Z}_{2}. We first construct some U𝒮iΛntU\in\mathcal{S}^{\Lambda\mathrm{nt}}_{\operatorname{i}\mathbb{R}} with a zero index.

To this end, we invoke Lemma˜B.5 separately on kerΛ,imΛ\ker\Lambda,\operatorname{im}\Lambda to obtain an ONB for \mathcal{H} {φi,ψi}i\Set{\varphi_{i},\psi_{i}}_{i\in\mathbb{Z}} such that i0i\leq 0 spans the kernel and i>0i>0 spans the image, and moreover, Cφi=ψiC\varphi_{i}=\psi_{i} for ii\in\mathbb{Z}. Let now Pφ±,Pψ±P^{\pm}_{\varphi},P^{\pm}_{\psi} be self-adjoint projections onto {φi}i>0,{ψi}i>0\Set{\varphi_{i}}_{i>0},\Set{\psi_{i}}_{i>0} and {φi}i0,{ψi}i0\Set{\varphi_{i}}_{i\leq 0},\Set{\psi_{i}}_{i\leq 0} respectively, and define

U=(PφPψ)(Pφ+Pψ+).\displaystyle U=(P_{\varphi}^{-}-P_{\psi}^{-})\oplus(P_{\varphi}^{+}-P_{\psi}^{+})\,.

Then UU is a Λ\Lambda-non-trivial SAU that belongs to i.\mathcal{B}_{\operatorname{i}\mathbb{R}}.

We turn to the construction of a U𝒮iΛntU\in\mathcal{S}^{\Lambda\mathrm{nt}}_{\operatorname{i}\mathbb{R}} that has a non-trivial index. Using Lemma˜B.5 separately on kerΛ,imΛ\ker\Lambda,\operatorname{im}\Lambda, we have an orthonormal basis fixed by CC for each of these spaces. Pick out a vector η±\eta^{\pm} out of each, and re-apply the lemma on kerΛspan(η),imΛspan(η+)\ker\Lambda\ominus\operatorname{span}(\eta^{-}),\operatorname{im}\Lambda\ominus\operatorname{span}(\eta^{+}) to obtain {φi,ψi}i\Set{\varphi_{i},\psi_{i}}_{i\in\mathbb{Z}} such that i0i\leq 0 spans the kernel minus η\eta^{-} and i>0i>0 spans the image minus η+\eta^{+}, and moreover, Cφi=ψiC\varphi_{i}=\psi_{i} for ii\in\mathbb{Z}. We define Pφ±,Pψ±P^{\pm}_{\varphi},P^{\pm}_{\psi} similarly as in the previous paragraph and define

U=[PφPψiη(η+)iη+(η)Pφ+Pψ+]\displaystyle U=\begin{bmatrix}P_{\varphi}^{-}-P_{\psi}^{-}&-\operatorname{i}\eta^{-}\otimes\left(\eta^{+}\right)^{\ast}\\ \operatorname{i}\eta^{+}\otimes\left(\eta^{-}\right)^{\ast}&P_{\varphi}^{+}-P_{\psi}^{+}\end{bmatrix}

Then U𝒮iΛntU\in\mathcal{S}^{\Lambda\mathrm{nt}}_{\operatorname{i}\mathbb{R}} and indΛ,2(U)=dimker(Pφ+Pψ+)=1\operatorname{ind}_{\Lambda,2}(U)=\dim\ker(P_{\varphi}^{+}-P_{\psi}^{+})=1 as ker(Pφ+Pψ+)=span(η+)\ker(P_{\varphi}^{+}-P_{\psi}^{+})=\operatorname{span}(\eta^{+}).

Let us now establish injectivity. Let U,U~𝒮𝔽ΛntU,\widetilde{U}\in\mathcal{S}^{\Lambda\mathrm{nt}}_{\mathbb{F}} be given, with 𝔽=i,i\mathbb{F}=\operatorname{i}\mathbb{R},\operatorname{i}\mathbb{H} and FF accordingly, and if 𝔽=i\mathbb{F}=\operatorname{i}\mathbb{R}, we assume that both operators have zero index (the non-trivial will be dealt with later on). We will deform UU~U\rightsquigarrow\widetilde{U} by a similar procedure to Theorem˜4.2: first we deform UU to a Λ\Lambda-diagonal SAU VV, and then we deform VV~V\rightsquigarrow\widetilde{V}. So now we write UU in eq.˜4.4 and we will reuse the properties proven and the construction defined in the proof of Theorem˜4.2. Since FF is hyper-local, {U,F}=0\Set{U,F}=0 implies that

(4.16) X,Y,A𝔽.\displaystyle X,Y,A\in\mathcal{B}_{\mathbb{F}}\,.

We would have liked to use Vf+(X)f(Y)V\equiv f_{+}(X)\oplus f_{-}(Y) from eq.˜4.11; that construction, however, is unsatisfactory since now VV does not respect the 𝔽\mathbb{F} symmetry constraint:

f+(X)F=(sgn(X)+χ{0}(X))F=F(sgn(X)+χ{0}(X))=Ff(X)Ff+(X).\displaystyle f_{+}(X)F=(\operatorname{sgn}(X)+\chi_{\Set{0}}(X))F=F(-\operatorname{sgn}(X)+\chi_{\Set{0}}(X))=-Ff_{-}(X)\neq-Ff_{+}(X)\,.

To fix this, we will decompose the zero eigenspace of XX into two disjoint parts of the same dimension χ{0}(X)=E+L\chi_{\Set{0}}(X)=E+L such that

(4.17) EF=FL.\displaystyle EF=FL\,.

To do so, we should have kerX\ker X even dimensional. In the case 𝔽=i\mathbb{F}=\operatorname{i}\mathbb{H}, the kernel is always even dimensional. Indeed, since Xφ=0X\varphi=0 iff XFφ=FXφ=0XF\varphi=-FX\varphi=0, it follows that χ{0}(X)F=Fχ{0}(X)\chi_{\Set{0}}(X)F=F\chi_{\Set{0}}(X). Hence F=JF=J is a symmetry operator on imχ{0}(X)kerX\operatorname{im}\chi_{\Set{0}}(X)\equiv\ker X and hence dimkerX2\dim\ker X\in 2\mathbb{N} from Lemma˜B.5, and we get an ONB of Kramers pairs {φi,ψi}i=1m\Set{\varphi_{i},\psi_{i}}_{i=1}^{m} for kerX\ker X. Let

(4.18) E:=i=1mφiφi,L:=i=1mψiψi.\displaystyle E:=\sum_{i=1}^{m}\varphi_{i}\otimes\varphi_{i}^{*},\quad L:=\sum_{i=1}^{m}\psi_{i}\otimes\psi_{i}^{*}\,.

Then

LJξ=i=1mψi,Jξψi=i=1mξ,Jψi,ψi=i=1mξ,φi,Jφi=i=1mJφi,ξi,φi=JEξ.\displaystyle LJ\xi=\sum_{i=1}^{m}\langle\psi_{i},J\xi\rangle\psi_{i}=\sum_{i=1}^{m}\langle\xi,J^{*}\psi_{i},\psi\rangle_{i}=\sum_{i=1}^{m}\langle\xi,\varphi_{i},J\rangle\varphi_{i}=\sum_{i=1}^{m}J\langle\varphi_{i},\xi_{i},\varphi\rangle_{i}=JE\xi\,.

For the case 𝔽=i\mathbb{F}=\operatorname{i}\mathbb{R} we need to further impose that dimkerX2\dim\ker X\in 2\mathbb{N} (which is equivalent to indΛ,2(U)=0\operatorname{ind}_{\Lambda,2}(U)=0, using the fact that the total operator is unitary, and stability of the index under compacts, see ˜A.4)–we deal with the odd case in the end. Hence for 𝔽=i\mathbb{F}=\operatorname{i}\mathbb{R} and F=CF=C, apply Lemma˜B.5 on kerX\ker X to obtain an ONB {φi,ψi}i=1m\Set{\varphi_{i},\psi_{i}}_{i=1}^{m} for with Cφi=ψiC\varphi_{i}=\psi_{i}. Now we define similarly EE and LL as in eq.˜4.18.

Note that eq.˜4.8 implies χ{0}(Y)=A1χ{0}(X)A\chi_{\Set{0}}(Y)=A^{-1}\chi_{\Set{0}}(X)A (where we mean AA as the isomorphism in eq.˜4.6). In fact, using eq.˜4.7 with λ=0\lambda=0 we see that AA maps kerY\ker Y unitarily onto kerX\ker X. Thus we can write χ{0}(Y)=Aχ{0}(X)A\chi_{\Set{0}}(Y)=A^{*}\chi_{\Set{0}}(X)A. Let us therefore define the diagonal SAU that does satisfy the symmetry constraint

V=[sgn(X)+EL00sgn(Y)A(EL)A]=:VLVR.\displaystyle V=\begin{bmatrix}\operatorname{sgn}(X)+E-L&0\\ 0&\operatorname{sgn}(Y)-A^{*}(E-L)A\end{bmatrix}=:V_{L}\oplus V_{R}\,.

Indeed, V𝔽V\in\mathcal{B}_{\mathbb{F}} since

VLF=(sgn(X)+EL)F=F(sgn(X)+LE)=FVL\displaystyle V_{L}F=(\operatorname{sgn}(X)+E-L)F=F(-\operatorname{sgn}(X)+L-E)=-FV_{L}

and similarly for VRV_{R}. Moreover, VV is a SAU since a short calculation (which uses the intertwining property eq.˜4.5) shows VL2=𝟙Λ,VR2=𝟙ΛV_{L}^{2}=\mathds{1}_{\Lambda^{\perp}},V_{R}^{2}=\mathds{1}_{\Lambda}.

Now that we have a diagonal symmetric VV, we may define the conjugation operator GG as before in eq.˜4.12. One has to be slightly careful since the different definition of VV leads to a different GG compared with eq.˜4.12, however, the two differ by a compact. Hence, all properties of GG from before still hold, and in particular, it is Fredholm of zero index. Similarly we can show that GG is invertible, and again it suffices to check that kerG={0}\ker G=\Set{0} since indG=0\operatorname{ind}G=0. To that end, suppose G[φψ]=0G\begin{bmatrix}\varphi\\ \psi\end{bmatrix}=0, then

(X+VL)φ+Aψ=0\displaystyle(X+V_{L})\varphi+A\psi=0
Aφ+(Y+VR)ψ=0.\displaystyle A^{*}\varphi+(Y+V_{R})\psi=0\,.

Using the intertwining property eq.˜4.5, we have

A(X+VL)=A(X+sgn(X)+EL)=(Ysgn(Y)+A(EL)A)A=(Y+VR)A.\displaystyle A^{*}(X+V_{L})=A^{*}(X+\operatorname{sgn}(X)+E-L)=(-Y-\operatorname{sgn}(Y)+A^{*}(E-L)A)A^{*}=-(Y+V_{R})A^{*}\,.

Hence when we multiply the first equation by AA^{*} we get

0=A(X+VL)φ+AAψ=(Y+VR)Aφ+|A|2ψ=((Y+VR)2+𝟙Y2)ψ\displaystyle 0=A^{*}(X+V_{L})\varphi+A^{*}A\psi=-(Y+V_{R})A^{*}\varphi+|A|^{2}\psi=((Y+V_{R})^{2}+\mathds{1}-Y^{2})\psi

where in the last step we use the second equation and |A|2=𝟙Y2.|A|^{2}=\mathds{1}-Y^{2}. Now

(Y+VR)2+𝟙Y2=Y2+VR2+YVR+VRY+𝟙Y2=2𝟙+YVR+VRY.\displaystyle(Y+V_{R})^{2}+\mathds{1}-Y^{2}=Y^{2}+V_{R}^{2}+YV_{R}+V_{R}Y+\mathds{1}-Y^{2}=2\mathds{1}+YV_{R}+V_{R}Y\,.

But observe that

YVR=Y(sgn(Y)A(EL)A)=Ysgn(Y)+AX(EL)=0A=Ysgn(Y)=|Y|.\displaystyle YV_{R}=Y(\operatorname{sgn}(Y)-A^{*}(E-L)A)=Y\operatorname{sgn}(Y)+A^{*}\underbrace{X(E-L)}_{=0}A=Y\operatorname{sgn}(Y)=|Y|\,.

Thus (Y+VR)2+𝟙Y2=2𝟙+2|Y|(Y+V_{R})^{2}+\mathds{1}-Y^{2}=2\mathds{1}+2|Y|, which implies that ψ=0\psi=0. Similarly, one can show that φ=0\varphi=0.

Since U,V𝔽U,V\in\mathcal{B}_{\mathbb{F}}^{\mathcal{L}}, it follows that G𝔽G\in\mathcal{B}_{\mathbb{F}}^{\mathcal{L}}, and hence W:=pol(G)𝒰𝔽W:=\operatorname{pol}(G)\in\mathcal{U}^{\mathcal{L}}_{\mathbb{F}}, and we have V=WUWV=W^{*}UW similar to eq.˜4.15. The rest of the arguments follow analogously to the proof in Theorem˜4.2, with the exception, however of an equivariant version (adapted to i\operatorname{i}\mathbb{R} and i\operatorname{i}\mathbb{H}) of Theorem˜B.1. To that end, let us rather apply Theorem˜3.1 on iW-\operatorname{i}W. Indeed, since

(iW)F=i(FW)=F(iW)\displaystyle(-\operatorname{i}W)F=-\operatorname{i}(-FW)=F(-\operatorname{i}W)

it follows that iW𝒰i𝔽-\operatorname{i}W\in\mathcal{U}^{\mathcal{L}}_{-\operatorname{i}\mathbb{F}} with 𝔽=i,i\mathbb{F}=\operatorname{i}\mathbb{R},\operatorname{i}\mathbb{H}, i.e., iW-\operatorname{i}W is a standard real or quaternionic operator, and there is a path from iW-\operatorname{i}W to 𝟙\mathds{1} within 𝒰i𝔽\mathcal{U}^{\mathcal{L}}_{-\operatorname{i}\mathbb{F}}. We multiply by i\operatorname{i} again to obtain a path from WW to i𝟙\operatorname{i}\mathds{1} within 𝒰𝔽.\mathcal{U}^{\mathcal{L}}_{\mathbb{F}}. Thus VV can be deformed to (i𝟙)U(i𝟙)=U(\operatorname{i}\mathds{1})^{*}U(\operatorname{i}\mathds{1})=U in 𝒮𝔽Λnt.\mathcal{S}^{\Lambda\mathrm{nt}}_{\mathbb{F}}.

Finally, we tackle the problem of connecting two operators U,U~𝒮iΛntU,\widetilde{U}\in\mathcal{S}^{\Lambda\mathrm{nt}}_{\operatorname{i}\mathbb{R}} both of whom have non-zero index. However, we can’t exactly follow the strategy above since it turns out that it is not possible to deform UU to a diagonal SAU that obeys that symmetry constraint. Since, by definition, ind2,ΛU=1\operatorname{ind}_{2,\Lambda}U=1 means dimkerX2+1\dim\ker X\in 2\mathbb{N}+1 (still using eq.˜4.4), we decompose this kernel into three disjoint parts

χ{0}(X)=E+L+ηη\displaystyle\chi_{\Set{0}}(X)=E+L+\eta\otimes\eta^{*}

where E,LE,L satisfies eq.˜4.17, and η(imΛ)\eta\in(\operatorname{im}\Lambda)^{\perp} is fixed by the symmetry Cη=ηC\eta=\eta. It is possible to find such η.\eta. Indeed, apply Lemma˜B.5 to construct an ONB of kerX\ker X fixed by CC, and pick one η\eta in this collection. Now apply Lemma˜B.5 again on kerXη\ker X\ominus\mathbb{C}\eta, and obtain {φi,ψi}i=1m\Set{\varphi_{i},\psi_{i}}_{i=1}^{m} such that Cφi=ψiC\varphi_{i}=\psi_{i}, and we construct E,LE,L similar as before eq.˜4.18. Define

(4.19) ξ:=AηimΛ\displaystyle\xi:=A^{*}\eta\in\operatorname{im}\Lambda

so that Cξ=CAη=ACη=Aη=ξC\xi=CA^{*}\eta=-A^{*}C\eta=-A^{*}\eta=-\xi, where we use eq.˜4.16 to show that CA=ACCA^{*}=-A^{*}C. Construct a SAU

(4.20) V:=[sgn(X)+ELηξξηsgn(Y)A(EL)A]=:[VLηξξηVR].\displaystyle V:=\begin{bmatrix}\operatorname{sgn}(X)+E-L&\eta\otimes\xi^{*}\\ \xi\otimes\eta^{*}&\operatorname{sgn}(Y)-A^{*}(E-L)A\end{bmatrix}=:\begin{bmatrix}V_{L}&\eta\otimes\xi^{*}\\ \xi\otimes\eta^{*}&V_{R}\end{bmatrix}\,.

We have V2=𝟙V^{2}=\mathds{1} from the following computation:

(sgn(X)+EL)2+(ηξ)(ξη)=sgn(X)2+(E+L+ηη)=(sgn(X)+χ{0}(X))2=𝟙\displaystyle(\operatorname{sgn}(X)+E-L)^{2}+(\eta\otimes\xi^{*})(\xi\otimes\eta^{*})=\operatorname{sgn}(X)^{2}+(E+L+\eta\otimes\eta^{*})=(\operatorname{sgn}(X)+\chi_{\Set{0}}(X))^{2}=\mathds{1}

and

(ξη)(sgn(X)+EL)+(sgn(Y)A(EL)A)(ξη)=0.\displaystyle(\xi\otimes\eta^{*})(\operatorname{sgn}(X)+E-L)+(\operatorname{sgn}(Y)-A^{*}(E-L)A)(\xi\otimes\eta^{*})=0\,.

To verify that ViV\in\mathcal{B}_{\operatorname{i}\mathbb{R}}, we compute

VC\displaystyle VC =[sgn(X)+ELηξξηsgn(Y)A(EL)A][C00C]\displaystyle=\begin{bmatrix}\operatorname{sgn}(X)+E-L&\eta\otimes\xi^{*}\\ \xi\otimes\eta^{*}&\operatorname{sgn}(Y)-A^{*}(E-L)A\end{bmatrix}\begin{bmatrix}C&0\\ 0&C\end{bmatrix}
=[(sgn(X)+EL)C(ηξ)C(ξη)C(sgn(Y)A(EL)A)C]\displaystyle=\begin{bmatrix}(\operatorname{sgn}(X)+E-L)C&(\eta\otimes\xi^{*})C\\ (\xi\otimes\eta^{*})C&(\operatorname{sgn}(Y)-A^{*}(E-L)A)C\end{bmatrix}
=[C(sgn(X)+LE)C(ηξ)C(ξη)C(sgn(Y)A(LE)A)]\displaystyle=\begin{bmatrix}C(-\operatorname{sgn}(X)+L-E)&-C(\eta\otimes\xi^{*})\\ -C(\xi\otimes\eta^{*})&C(-\operatorname{sgn}(Y)-A^{*}(L-E)A)\end{bmatrix}
=CV\displaystyle=-CV

where we use the fact that AC=CAAC=-CA and Cη=ηC\eta=\eta. Now that we have an appropriate SAU VV, we follow eq.˜4.12 and define G:=12(U+V)G:=\frac{1}{2}(U+V). This self-adjoint operator enjoys all the properties discussed above of having indΛG=0\operatorname{ind}_{\Lambda}G=0 and invertible. To see the invertibility, one follows a similar calculation to the one performed already twice above, so we omit it here.

Using the invertibility of GG we may construct now a path from UVU\rightsquigarrow V within 𝒮iΛnt\mathcal{S}^{\Lambda\mathrm{nt}}_{\operatorname{i}\mathbb{R}}. Having deformed U,U~U,\widetilde{U} into V,V~V,\widetilde{V} respectively we now seek a unitary operator that conjugates VV into V~\widetilde{V}, and could be deformed to 𝟙\mathds{1} in a certain symmetric way. Using the RHS of eq.˜4.20, decompose the space kerΛ\ker\Lambda as

kerΛ=imχ{1}(VL)imχ{1}(VL)χ{0}(VL)=:𝒢\displaystyle\ker\Lambda=\operatorname{im}\chi_{\Set{1}}(V_{L})\oplus\operatorname{im}\chi_{\Set{-1}}(V_{L})\oplus\chi_{\Set{0}}(V_{L})=:\mathcal{G}

where, in fact, imχ{0}(VL)=span(η)\operatorname{im}\chi_{\Set{0}}(V_{L})=\operatorname{span}(\eta). We may decompose kerΛ\ker\Lambda similarly according to V~L\widetilde{V}_{L}, and we denote this grading as 𝒢~\widetilde{\mathcal{G}}; note 𝒢\mathcal{G} and 𝒢~\widetilde{\mathcal{G}} are the same space kerΛ\ker\Lambda which is merely graded differently. Since VLC=CVLV_{L}C=-CV_{L}, it follows that χ{1}(VL)C=Cχ{1}(VL)\chi_{\Set{1}}(V_{L})C=C\chi_{\Set{-1}}(V_{L}). Since χ{1}(VL),χ{1}(VL~)\chi_{\Set{1}}(V_{L}),\chi_{\Set{1}}(\widetilde{V_{L}}) are both infinite dimensional, there is a unitary Z:χ{1}(VL)χ{1}(VL~)Z:\chi_{\Set{1}}(V_{L})\to\chi_{\Set{1}}(\widetilde{V_{L}}). Using it, we define WL:𝒢𝒢~W_{L}:\mathcal{G}\to\widetilde{\mathcal{G}} as

WL:=[Z000CZC000η~η].\displaystyle W_{L}:=\begin{bmatrix}Z&0&0\\ 0&CZC&0\\ 0&0&\widetilde{\eta}\otimes\eta^{\ast}\end{bmatrix}\,.

It is clear that WLW_{L} is unitary. In this grading, we can write C,VL:𝒢𝒢C,V_{L}:\mathcal{G}\to\mathcal{G} as

C=[0Cimχ{1}(VL)imχ{1}(VL)0Cimχ{1}(VL)imχ{1}(VL)0000𝟙],VL=[𝟙000𝟙0000].\displaystyle C=\begin{bmatrix}0&C_{\operatorname{im}\chi_{\Set{-1}}(V_{L})\to\operatorname{im}\chi_{\Set{1}}(V_{L})}&0\\ C_{\operatorname{im}\chi_{\Set{1}}(V_{L})\to\operatorname{im}\chi_{\Set{-1}}(V_{L})}&0&0\\ 0&0&\mathds{1}\end{bmatrix},\quad V_{L}=\begin{bmatrix}\mathds{1}&0&0\\ 0&-\mathds{1}&0\\ 0&0&0\end{bmatrix}\,.

A similar expression holds for 𝒢~\widetilde{\mathcal{G}} and VL~,C:𝒢~𝒢~\widetilde{V_{L}},C:\widetilde{\mathcal{G}}\to\widetilde{\mathcal{G}}. Now a direct computation shows that

VL=WLV~LWL,WLC=CWL.\displaystyle V_{L}=W_{L}^{\ast}\widetilde{V}_{L}W_{L},\quad W_{L}C=CW_{L}\,.

We define WR:imΛimΛW_{R}:\operatorname{im}\Lambda\to\operatorname{im}\Lambda similarly. Plugging this into the RHS of eq.˜4.20, we find

V\displaystyle V =[WLV~LWLηξξηWRV~RWR]\displaystyle=\begin{bmatrix}W_{L}^{\ast}\widetilde{V}_{L}W_{L}&\eta\otimes\xi^{*}\\ \xi\otimes\eta^{*}&W_{R}^{\ast}\widetilde{V}_{R}W_{R}\end{bmatrix}
=[WL00WR][V~LWLηξWRWRξηWLV~R][WL00WR].\displaystyle=\begin{bmatrix}W_{L}^{*}&0\\ 0&W_{R}^{*}\end{bmatrix}\begin{bmatrix}\widetilde{V}_{L}&W_{L}\eta\otimes\xi^{*}W_{R}^{*}\\ W_{R}\xi\otimes\eta^{*}W_{L}^{*}&\widetilde{V}_{R}\end{bmatrix}\begin{bmatrix}W_{L}&0\\ 0&W_{R}\end{bmatrix}\,.

In fact, it holds that WL(ηξ)WR=η~ξ~W_{L}(\eta\otimes\xi^{*})W_{R}^{*}=\widetilde{\eta}\otimes\widetilde{\xi}^{*}. Thus V=WV~WV=W\widetilde{V}W^{*} where W:=WLWRW:=W_{L}\oplus W_{R}.

We can use Theorem˜B.1 to deform each WL,WRW_{L},W_{R} to 𝟙\mathds{1} such that the path commutes with CC, which yields a path connecting VV~V\rightsquigarrow\widetilde{V} within 𝒮iΛnt\mathcal{S}^{\Lambda\mathrm{nt}}_{\operatorname{i}\mathbb{R}}. ∎

Lemma 4.4 (Λ\Lambda-non-triviality is well-defined).

Let U𝒮ΛntU\in\mathcal{S}^{\Lambda\mathrm{nt}} and W𝒮W\in\mathcal{S}^{\mathcal{L}}. If UWU-W is compact or sufficiently small in norm, then W𝒮ΛntW\in\mathcal{S}^{\Lambda\mathrm{nt}} too. In particular, any continuous path in 𝒮\mathcal{S}^{\mathcal{L}} starting within 𝒮Λnt\mathcal{S}^{\Lambda\mathrm{nt}} is entirely contained within 𝒮Λnt\mathcal{S}^{\Lambda\mathrm{nt}}.

Proof.

Let U𝒮ΛntU\in\mathcal{S}^{\Lambda\mathrm{nt}}. Then there exists VV as in ˜4.1 such that [V,Λ]=0[V,\Lambda]=0, and VV is non-trivial when restricted to either imΛ\operatorname{im}\Lambda or (imΛ)(\operatorname{im}\Lambda)^{\perp}, and UV𝒦U-V\in\mathcal{K}. Let W𝒮W\in\mathcal{S}^{\mathcal{L}}. Decompose WW in =(imΛ)imΛ\mathcal{H}=(\operatorname{im}\Lambda)^{\perp}\oplus\operatorname{im}\Lambda as

[WLLWLRWRLWRR].\displaystyle\begin{bmatrix}W_{LL}&W_{LR}\\ W_{RL}&W_{RR}\end{bmatrix}\,.

Here WiiW_{ii} is self-adjoint and essentially unitary for i=L,Ri=L,R. By Lemma˜2.3 and the fact that WiiUiiWU\left\lVert W_{ii}-U_{ii}\right\rVert\leq\left\lVert W-U\right\rVert for i=L,Ri=L,R, we conclude that WiiW_{ii} is essentially a non-trivial SAU.

The compact statement is trivial. ∎

5 Classification of bulk one-dimensional spectrally-gapped insulators

We now come to the classification of one-dimensional insulators with a spectral gap. Let us begin with the general setup. We are interested in describing quantum mechanical systems of non-interacting electrons on a lattice, and hence we choose the Hilbert space

:=2(d)N\mathcal{H}:=\ell^{2}(\mathbb{Z}^{d})\otimes\mathbb{C}^{N}

where dd is the space dimension and NN is the (fixed) number of internal degrees of freedom on each lattice site. The choice of the cubic lattice is made for simplicity of notation, since changing NN we may encode any graph via redimerization. What is however of importance is the fact d\mathbb{Z}^{d} has no boundary, which corresponds physically to bulk systems. Later we comment briefly on edge systems in Section˜6. We note that a classification of continuum systems with Hilbert space L2(d)NL^{2}(\mathbb{R}^{d})\otimes\mathbb{C}^{N} would also be interesting, especially since some of the features presented here seem to only emerge in the tight-binding setting, see [SW22a, SW22].

As was mentioned already above, locality plays a crucial role in our analysis. Physically it corresponds to the decaying probability of quantum mechanical transition between farther and farther points in space. There are various ways to encode locality of a quantum mechanical operator; Let {δx}xd\Set{\delta_{x}}_{x\in\mathbb{Z}^{d}} be the singled-out position basis of Hilbert space, so that for any AA\in\mathcal{B} and x,ydx,y\in\mathbb{Z}^{d}, the expression AxyA_{xy} corresponds to an N×NN\times N matrix whose matrix elements are

(Axy)ijδxei,Aδyej(i,j{1,,N})(A_{xy})_{ij}\equiv\langle\delta_{x}\otimes e_{i},A\delta_{y}\otimes e_{j}\rangle\qquad(i,j\in\Set{1,\dots,N})

with {ei}i=1N\Set{e_{i}}_{i=1}^{N} the standard basis for N\mathbb{C}^{N}. Now, the most straight forward way which is common in physics to specify locality is the nearest-neighbor constraint, i.e., AA\in\mathcal{B} is local iff

Axy=Axyχ{0,1}(xy)(x,yd)A_{xy}=A_{xy}\chi_{\Set{0,1}}(\left\lVert x-y\right\rVert)\qquad(x,y\in\mathbb{Z}^{d})

where we take, say, the Euclidean norm on d\mathbb{Z}^{d} and χ\chi is the characteristic function. Sometimes one prefers to consider finite hopping operators, which are those operators AA\in\mathcal{B} for which there exists some R>0R>0 such that

Axy=AxyχR(xy)(x,yd).\displaystyle A_{xy}=A_{xy}\chi_{\mathbb{R}_{\leq R}}(\left\lVert x-y\right\rVert)\qquad(x,y\in\mathbb{Z}^{d})\,.

In mathematics it is customary to consider the locality constraint as exponential decay of the off-diagonal matrix elements, i.e., that there exists some C,μ<C,\mu<\infty such that

(5.1) AxyCexp(μxy)(x,yd).\displaystyle\left\lVert A_{xy}\right\rVert\leq C\exp\left(-\mu\left\lVert x-y\right\rVert\right)\qquad(x,y\in\mathbb{Z}^{d})\,.

Here we may choose any matrix norm for the LHS. This definition of locality is very natural and also facilitates the analysis on many occasions, it has appeared in various papers on topological insulators, e.g. [EGS05, Sha20, Fon+20, BSS23, ST19].

In choosing the correct definition of locality there is a certain art. If we were to insist on the above definition via exponential decay eq.˜5.1, the analysis becomes tedious and inelegant. Indeed, to drive this point further, and out of general interest, we explore this idea later in Section˜5.6. On the other hand one may define locality as that property of operators so that (together with the gap condition), topological indices are well-defined, which might lead to rather abstract topological analysis. Here we choose a middle ground which on the one hand leads to relatively natural functional analytic proofs and on the other hand is somewhat of a shadow of eq.˜5.1. We formulate it only in one and two dimensions here so as to avoid additional notational overhead which is anyway not necessary in the present paper, but see Section˜8 below for the construction in higher dimensions.

Definition 5.1 (locality in d=1d=1).

Define an operator to be local iff it is Λ\Lambda-local as in ˜2.5, now with the particular choice Λ:=χ(X)\Lambda:=\chi_{\mathbb{N}}(X) where XX is the position operator on 2()\ell^{2}(\mathbb{Z}). Hence, AA\in\mathcal{B} is local iff [A,Λ]𝒦[A,\Lambda]\in\mathcal{K}.

Definition 5.2 (locality in d=2d=2).

Let X1,X2X_{1},X_{2} be the two position operators on 2()\ell^{2}(\mathbb{Z}), with which Φ=arg(X1+iX2)\Phi=\arg(X_{1}+\operatorname{i}X_{2}) is the angle-position operator and eiΦ\operatorname{e}^{\operatorname{i}\Phi} is the phase position operator. An operator AA\in\mathcal{B} is termed local now iff [eiΦ,A]𝒦[\operatorname{e}^{\operatorname{i}\Phi},A]\in\mathcal{K}.

It is a fact that eq.˜5.1 implies the compact commutator locality criterion: indeed, this is proven e.g. in [GS18, Lemma 2 (b)] and [BSS23, Lemma A.1] for d=1,2d=1,2 respectively. On the other hand it is certainly clear that these compact commutator notions of locality are strictly weaker than eq.˜5.1. From now on in this section \mathcal{L} stands for local operators with the compact commutator condition (very soon we will specify to d=1d=1 and then we mean ˜5.1).

Remark 5.3 (Compact commutator locality and the role of NN).

In our presentation so far the parameter NN is the internal fiber dimension, which physically could stand for spin, isospin, sub-lattice, or any other on-site internal degree of freedom of electrons. By requiring that operators are local via ˜5.1 instead of eq.˜5.1, we in principle allow NN to vary as we perform homotopies between operators. Indeed, by re-dimerization, given any operator presented on a Hilbert space with one given NN we may obtain another operator with any other N~\widetilde{N} and clearly both would obey the compact commutator condition. This is thus a counter point of criticism on our K-theoretic-free analysis: why go through so much trouble to avoid K-theory if in the end anyway NN may effectively vary during homotopies? One response would be that unlike in K-theory our construction still calculates absolute rather than relative phases (we avoid the Grothendiek construction) and moreover, as explained, the calculation brings the topology defined on the set of operators to the foreground and as such may allow us to deal with the mobility gap regime.

Definition 5.4 (material).

A material is then specified as a local quantum mechanical Hamiltonian HH on \mathcal{H}, i.e., some self-adjoint bounded linear operator H=HH=H^{\ast}\in\mathcal{L}.

5.1 Insulators

The space of all materials is too big to be topologically interesting (it is clearly nullhomotopic with straight-line homotopies). To further restrict it, we concentrate on insulators: materials which exhibit zero direct current if electric voltage is applied. This statement needs to be qualified: due to the Pauli exclusion principle, electrons in a solid are characterized by a Fermi energy μ\mu\in\mathbb{R}, and so the same material could be both an insulator and a conductor when probed at different values of μ\mu. It turns out that for the purpose of conductivity, at a given μ\mu, it is equivalent to consider either HH at Fermi energy μ\mu, or Hμ𝟙H-\mu\mathds{1} at Fermi energy 0; clearly the latter operator is local too. Hence for the sake of simplicity we shall henceforth assume, without loss of generality, that the Fermi energy is always fixed at μ=0\mu=0. We note in passing that this assumption is not entirely benign when coupled with symmetries: further below we will see that certain symmetric operators have spectral symmetry about zero and then if one sets the Fermi energy at values other than zero one may obtain a different classification.

We identify two ways to encode the insulator (at μ=0\mu=0) condition: the spectral gap and the mobility gap criterions. The spectral gap condition is a simple constraint on the operator 0σ(H){0\notin\sigma(H)}, i.e., HH is an invertible operator. Since σ(H)Closed()\sigma(H)\in\mathrm{Closed}(\mathbb{R}), this implies the existence of an open interval about zero which is not in the spectrum. The mobility gap condition is rather a constraint on the quantum dynamics associated with HH, and is a set of almost-sure consequences for random ensembles of operators exhibiting Anderson localization. This condition was first presented in [EGS05]. Since we will discuss the mobility gap regime specifically later in Section˜7 let us continue with the general progression here and accept that insulators are

Definition 5.5 (insulators).

A material H=HH=H^{\ast}\in\mathcal{L} is an insulator iff it is invertible, i.e., if 0σ(H)0\notin\sigma(H). The space of all insulators is denoted by N\mathcal{I}\equiv\mathcal{I}_{N} (we mostly keep the fiber dimension NN implicit since it is fixed) and is endowed, as all other spaces, with the subspace topology from the operator norm topology on \mathcal{B}.

To each insulator HH we associate a Fermi projection

PP(H):=χ(,0)(H)\displaystyle P\equiv P(H):=\chi_{(-\infty,0)}(H)

which physically speaking corresponds to the Fermionic many-body ground state (density matrix) within the single-particle Hilbert space. Importantly, PP inherits locality from HH: This is a consequence of Lemma˜2.8 and the fact that under the assumption of a spectral gap, χ(,0)(H)=f(H)\chi_{(-\infty,0)}(H)=f(H) with ff a continuous function differing from χ(,0)\chi_{(-\infty,0)} on σ(H)\mathbb{C}\setminus\sigma(H).

At this point we specify to d=1d=1. The task at hand is to calculate π0()\pi_{0}(\mathcal{I}). According to the Kitaev table Table˜1 we should recover π0()={0}\pi_{0}(\mathcal{I})=\Set{0}. This is however not true at the level of generality we are working. Indeed, this is clear even without locality constraints: just take any insulator that has spectrum only above zero and another insulator that has spectrum only below zero: these two cannot be connected without passing with spectrum through zero and hence exiting \mathcal{I}. A remedy would be to constrain to the space of insulators such that their Fermi projection is non-trivial as in ˜2.1. But actually even this is still not enough: locality in one-dimension divides the system into left and right halves, and we should insist that our system is non-trivial on each side separately–this is the notion of Λ\Lambda-non-trivial projections from ˜4.1 (adapted from SAUs to self-adjoint projections in an obvious way)–so that we are speaking about genuine bulk systems rather than domain walls or edge systems.

Example 5.6 (The necessity of Λ\Lambda-non-triviality).

Let H:=ΛΛH:=\Lambda-\Lambda^{\perp} and H~:=Λ+Λ\widetilde{H}:=-\Lambda+\Lambda^{\perp}. Both of these (flat) Hamiltonians are local (indeed, diagonal in space and in energy) and each has a Fermi projection which is non-trivial in the sense of ˜2.1, because it has an infinite kernel and infinite range. However, on each half of space separately, the Fermi projections are trivial (just 𝟙\mathds{1} or 0).

We claim that HH cannot be deformed into H~\widetilde{H} without either closing the gap or violating locality.

Proof.

We prove the claim by contradiction: suppose there exists a continuous path tHtt\mapsto H_{t} that deforms HH to H~\widetilde{H} such that HtH_{t} is self-adjoint, invertible and local. Then t12(𝟙sgn(Ht))t\mapsto\frac{1}{2}(\mathds{1}-\operatorname{sgn}(H_{t})) is a continuous path of local self-adjoint projections that connects the Fermi projection of HH to H~\widetilde{H}, which we denote as PP and P~\widetilde{P}, respectively. Let us recall [Rør+00, Proposition 2.2.6], which says that for any C-star algebra 𝒜\mathcal{A}, if A,B𝒜A,B\in\mathcal{A} are projections that are path-connected, then there exists a unitary in 𝒜\mathcal{A} conjugating them. We apply this lemma on the C-star algebra \mathcal{L} to conclude that there exists some U𝒰U\in\mathcal{U}^{\mathcal{L}} such that P=UP~UP=U^{*}\widetilde{P}U. Decompose UU in =(imΛ)imΛ\mathcal{H}=(\operatorname{im}\Lambda)^{\perp}\oplus\operatorname{im}\Lambda as eq.˜3.6. Writing out the equation P=UP~UP=U^{*}\widetilde{P}U in this decomposition, we find

[𝟙000]=[ULLURLULRURR][000𝟙][ULLULRURLURR]=[URLURLURLURRURRURLURRURR].\displaystyle\begin{bmatrix}\mathds{1}&0\\ 0&0\end{bmatrix}=\begin{bmatrix}U^{*}_{LL}&U^{*}_{RL}\\ U^{*}_{LR}&U^{*}_{RR}\end{bmatrix}\begin{bmatrix}0&0\\ 0&\mathds{1}\end{bmatrix}\begin{bmatrix}U_{LL}&U_{LR}\\ U_{RL}&U_{RR}\end{bmatrix}=\begin{bmatrix}U^{*}_{RL}U_{RL}&U^{*}_{RL}U_{RR}\\ U^{*}_{RR}U_{RL}&U^{*}_{RR}U_{RR}\end{bmatrix}\,.

Thus URLURL=𝟙.U_{RL}^{*}U_{RL}=\mathds{1}. Now by assumption UU is local, which implies that URLU_{RL} is a compact operator. However, 𝟙\mathds{1} is not compact on the infinite-dimensional space (imΛ)(\operatorname{im}\Lambda)^{\perp}. This leads to a contradiction. ∎

We thus define

Definition 5.7 (bulk insulators).

A material HH\in\mathcal{I} is a bulk-insulator iff its Fermi projection is Λ\Lambda-non-trivial, i.e., sgn(H)\operatorname{sgn}(H) is a Λ\Lambda-non-trivial SAU in the sense of ˜4.1. We denote the space of bulk-insulators with BN,B\mathcal{I}_{B}\equiv\mathcal{I}_{N,B}:

(5.2) B:={H|sgn(H)𝒮Λnt}\displaystyle\mathcal{I}_{B}:=\Set{H\in\mathcal{I}}{\operatorname{sgn}(H)\in\mathcal{S}^{\Lambda\mathrm{nt}}}

and furnish it also with the subspace topology.

It will indeed emerge that in one space dimension, π0(B)={0}\pi_{0}(\mathcal{I}_{B})=\Set{0} as stipulated by Table˜1; this is one case of Theorem˜5.12 below.

5.2 The Altland-Zirnbauer symmetry classes

Next we discuss the Altland-Zirnbauer symmetry classes [AZ97] (AZ classes henceforth). The idea is that by restricting to a subspace, we could obtain non-trivial topology. From context of physics, naturally the subspaces of operators are those which obey certain symmetries. According to Wigner’s theorem [Bar64], a symmetry is a unitary or anti-unitary operator on \mathcal{H}. Two basic operations coming from quantum field theory are time-reversal Θ\Theta and charge conjugation 𝒞\mathcal{C} (which, in the context of solid state physics should be considered as particle-hole Ξ\Xi); the third one is parity which we do not need here. Naturally since the time evolution in quantum mechanics is implemented via exp(itH)\exp(-\operatorname{i}tH), Θ\Theta should be anti-unitary and HH is deemed “time-reversal invariant” iff it commutes with Θ\Theta. It was Dyson [Dys62] who identified the two important cases Θ2=±𝟙\Theta^{2}=\pm\mathds{1} which eponymously became known as Dyson’s three-fold way (no Θ\Theta constraint or [H,Θ]=0[H,\Theta]=0 with Θ=±𝟙\Theta=\pm\mathds{1}). Altland and Zirnbauer [AZ97] combined the three-fold way together with the charge-conjugation operator to form what is now known as the ten-fold way. They considered many-body systems and Bogoliubov-de-Gennes (BdG) Hamiltonian description of superconductors, and in the context of which, one may think of particle hole Ξ\Xi again as an anti-unitary operator which may square to ±𝟙\pm\mathds{1}, and commutes with Θ\Theta. However, now, a Hamiltonian is deemed particle-hole symmetric iff it anti-commutes with Ξ\Xi:

(5.3) {H,Ξ}HΞ+ΞH=0.\displaystyle\{H,\Xi\}\equiv H\Xi+\Xi H=0\,.

The idea that a symmetry anti-commutes with a Hamiltonian may appear unnatural and at odds with basic notions of quantum mechanics–this is not how Altland and Zirnbauer phrased their many-body theory where all symmetries commute with the Hamiltonian; see [Zir21] for further discussion. Nonetheless it became quite established in modern condensed matter physics to use the anti-commutation condition as a convenient way to deal with particle-hole symmetry, and we will follow suit. They then defined the chiral symmetry operator as the composition of the two

Π:=ΘΞ.\displaystyle\Pi:=\Theta\Xi\,.

Since both Θ\Theta and Ξ\Xi are anti-unitary, Π\Pi is actually unitary and its square is of no consequence in the sense that {H,Π}=0\{H,\Pi\}=0 iff {H,iΠ}=0\{H,\operatorname{i}\Pi\}=0. An interesting point is that one may consider a system which is chiral-symmetric (so it obeys {H,Π}=0\{H,\Pi\}=0 even though it has no further symmetries). Taking into account all possibilities (presence or absence of each symmetry constraint with each ±𝟙\pm\mathds{1} version) we arrive at ten possibilities which are depicted in the first column of Table˜1. These ten possibilities correspond to well-known structures in mathematics, such as the ten Morita equivalence classes of Clifford algebras [ABS64], Cartan’s ten infinite families of compact symmetric spaces [Car26, Car27] and the ten associative real super division algebras [Wal64, Del99]. The AZ labels themselves, by the way, come from Cartan.

Assumption 5.8 (Symmetries are strictly local).

We shall assume that Θ,Ξ\Theta,\Xi and Π\Pi are strictly local, i.e., they commute with the position operator XX. Hence they can be considered as (anti-)unitary operators on N\mathbb{C}^{N}.

It would appear that most of the analysis should probably go through if it is only assumed that the commutator is compact: redimerization could make it hold if the symmetry operators have finite range.

Remark 5.9.

In the foregoing discussion, we merely remarked that the sign of Π2\Pi^{2} is of no consequence to the analysis, and usually, when one presents the Kitaev table Table˜1 (as we did) one does not write out what Π2\Pi^{2} is, but rather only whether it is present or not.

It is however clear that if Θ\Theta and Ξ\Xi are presumed to commute (as we indeed assume) then Π2=Θ2Ξ2\Pi^{2}=\Theta^{2}\Xi^{2} and hence according to Table˜1 once Θ2\Theta^{2} and Ξ2\Xi^{2} disagree, Π2=𝟙\Pi^{2}=-\mathds{1}. This however contradicts the ubiquitous convention of taking Π=𝟙σ3\Pi=\mathds{1}\otimes\sigma_{3} which always squares to +𝟙+\mathds{1}. Thus there are two possibilities: either take Π~=i𝟙σ3\widetilde{\Pi}=\operatorname{i}\mathds{1}\otimes\sigma_{3} for those AZ symmetry classes where Ξ2\Xi^{2} and Θ2\Theta^{2} disagree ({H,Π}=0\Set{H,\Pi}=0 iff {H,Π~}=0\Set{H,\widetilde{\Pi}}=0), or equivalently, for those AZ symmetry classes, take {Θ,Π}=0\Set{\Theta,\Pi}=0 instead of [Θ,Π]=0[\Theta,\Pi]=0.

To preserve notational simplicity, we found it more convenient to always assume that Π=𝟙σ3\Pi=\mathds{1}\otimes\sigma_{3} and when necessary, employ {Θ,Π}=0\Set{\Theta,\Pi}=0; this convention follows, e.g., [KK18]. This explains the following assumption.

Assumption 5.10.

We assume that Π\Pi has ±1\pm 1 eigenspaces of the same dimension, and that there is a unitary mapping between the two Π\Pi eigenspaces which commutes with both Θ\Theta or Ξ\Xi.

Definition 5.11 (symmetric insulators).

To each of the AZ symmetry classes

AZ:={A,AI,AII,AIII,BDI,D,DIII,C,CI,CII}\displaystyle\mathrm{AZ}:=\Set{\mathrm{A},\mathrm{AI},\mathrm{AII},\mathrm{AIII},\mathrm{BDI},\mathrm{D},\mathrm{DIII},\mathrm{C},\mathrm{CI},\mathrm{CII}}

we define the class of bulk-insulators which obey that symmetry and label it by

B,ΣN,B,Σ(ΣAZ).\displaystyle\mathcal{I}_{B,\Sigma}\equiv\mathcal{I}_{N,B,\Sigma}\qquad(\Sigma\in\mathrm{AZ})\,.

The main result of this section is

Theorem 5.12 (The one-dimensional column of the Kitaev table).

At each fixed NN, for any ΣAZ\Sigma\in\mathrm{AZ}, the path-connected components of N,B,Σ\mathcal{I}_{N,B,\Sigma} considered with the subspace topology associated with the operator norm topology, agree with the set appearing in the first column of Table˜1.

We stress that while Table˜1 was derived using K-theory of C-star algebras, here we make no recourse to K-theory and rely entirely on homotopies of operators. In particular, the classification we derive is not relative and does not rely on extended degrees of freedom (for us NN in =2()N\mathcal{H}=\ell^{2}(\mathbb{Z})\otimes\mathbb{C}^{N} is fixed once and for all throughout the analysis). While these two points might not exactly appeal to specialists in K-theory, what is perhaps more interesting is the perspective on the mobility gap regime, see Section˜7.

The rest of this section is dedicated to proving Theorem˜5.12 using the results presented in Sections˜3 and 4. In Section˜5.6 we present a completely different approach which assumes a different mode of locality via eq.˜5.1.

Examples of concrete physical models.

In order to connect with concrete literature in physics, we point out that

  1. 1.

    In class AIII, the Hamiltonian is of the form

    H=[0SS0]\displaystyle H=\begin{bmatrix}0&S^{\ast}\\ S&0\end{bmatrix}

    and the associated index is

    indΛpol(S).\displaystyle\operatorname{ind}\mathbb{\Lambda}\operatorname{pol}(S)\in\mathbb{Z}\,.

    This index is widely known as the Zak phase [Zak89], and in the translation-invariant setting reduces to a winding number. A widely popular model which exhibits a non-trivial Zak phase is the SSH model [SSH79] of polyacetylene.

  2. 2.

    In class D, the associated index is

    ind2Λsgn(H)2.\displaystyle\operatorname{ind}_{2}\mathbb{\Lambda}\operatorname{sgn}(H)\in\mathbb{Z}_{2}\,.

    This index is widely known as the Majorana number and a quintessential model which exhibits it is the Kitaev chain [Kit01].

5.3 Flat Hamiltonians

Sometimes in physics there is a distinction between classifying Hamiltonians and classifying ground states, which, in the single-particle context correspond to the associated Fermi projections. As we will see now, for us this distinction does not exist since we are working in the spectral gap regime.

We say a Hamiltonian HH is flat iff sgn(H)=H\operatorname{sgn}(H)=H where sgn\operatorname{sgn} is the sign function (its value at zero is of no consequence since our Hamiltonians have no spectrum there). We denote the space of all flat bulk-insulators by B\mathcal{I}_{B}^{\flat}. We note that if HH is flat then its Fermi projection PP is given by P=12(𝟙H)P=\frac{1}{2}\left(\mathds{1}-H\right) so flat Hamiltonians are algebraically related to their Fermi projections.

Lemma 5.13.

Flat insulators are a strong deformation retract of insulators. This statement remains true if we add the bulk-insulator constraint as well as any of the ten AZ symmetry constraints: B,Σ\mathcal{I}_{B,\Sigma}^{\flat} is a strong deformation retraction of B,Σ\mathcal{I}_{B,\Sigma} for any ΣAZ\Sigma\in\mathrm{AZ}.

Proof.

The desired retraction is in fact sgn\operatorname{sgn}, which (via the functional calculus) may be considered a map \mathcal{L}\to\mathcal{L} (see Lemma˜2.8).

Hence, given HB,ΣH\in\mathcal{I}_{B,\Sigma}, one has

sgn(H)=𝟙+1πiΓR(z)dz\displaystyle\operatorname{sgn}(H)=\mathds{1}+\frac{1}{\pi\operatorname{i}}\oint_{\Gamma}R(z)\operatorname{d}{z}

where Γ\Gamma is any CCW path encircling σ(H)(,0)\sigma(H)\cap(-\infty,0) and R(z)(Hz𝟙)1R(z)\equiv(H-z\mathds{1})^{-1}. From this formula and the resolvent identity norm continuity easily follows. Since sgnsgn=sgn\operatorname{sgn}\circ\operatorname{sgn}=\operatorname{sgn}, this is indeed a retraction; note that since sgn\operatorname{sgn} is odd, sgn(H)\operatorname{sgn}(H) would obey the same AZ constraint that HH would.

Next, define F:B,Σ×[0,1]B,ΣF:\mathcal{I}_{B,\Sigma}\times[0,1]\to\mathcal{I}_{B,\Sigma} via

F(H,t):=(1t)H+tsgn(H)(HB,Σ,t[0,1]).\displaystyle F(H,t):=(1-t)H+t\operatorname{sgn}(H)\qquad(H\in\mathcal{I}_{B,\Sigma},t\in[0,1])\,.

It is well-defined since

χ(,0)(F(t,H))=χ(,0)(H)(t[0,1])\displaystyle\chi_{(-\infty,0)}(F(t,H))=\chi_{(-\infty,0)}(H)\qquad(t\in[0,1])

and F(H,0)=HF(H,0)=H, F(H,1)=sgn(H)F(H,1)=\operatorname{sgn}(H) and F(sgn(H),1)=sgn(H)F(\operatorname{sgn}(H),1)=\operatorname{sgn}(H). Since the bulk-insulator condition is defined in terms of the flat Hamiltonian and not the Hamiltonian itself, F(t,H)F(t,H) is a bulk-insulator for all tt. ∎

Clearly the path-connected components of a space and those of its retract are the same, and hence in proving Theorem˜5.12, we could just as well work with B,Σ\mathcal{I}^{\flat}_{B,\Sigma}. This last fact makes the analysis reduce to the study of Λ\Lambda-non-trivial equivariant self-adjoint unitaries.

5.4 Classification of the non-chiral classes

𝔽\mathbb{F} AZ Class Topological invariant
\mathbb{C} A -
\mathbb{R} AI -
\mathbb{H} AII -
i\operatorname{i}\mathbb{R} D ind2Λsgn(H)\operatorname{ind}_{2}\mathbb{\Lambda}\operatorname{sgn}(H)
i\operatorname{i}\mathbb{H} C -
Table 4: Correspondence between the operators defined in eqs.˜2.6 and 2.8 and the non-chiral AZ symmetry classes classes, and formulas for the topological invariants.

The non-chiral classes are those within the AZ classes where Π\Pi is absent: classes A,AI,AII, C and D. In this case B,Σ\mathcal{I}_{B,\Sigma}^{\flat} is the same space as 𝒮𝔽Λnt\mathcal{S}^{\Lambda\mathrm{nt}}_{\mathbb{F}} with the appropriate correspondence between Σ\Sigma and 𝔽\mathbb{F} as depicted in Table˜4. When Θ\Theta squares to ±𝟙\pm\mathds{1}, we have a real (resp. quaternionic) structure and that corresponds to the anti-unitary operator CC (resp. JJ) of eq.˜2.6. On the other hand, the presence of a particle-hole symmetry corresponds rather to sgn(H)\operatorname{sgn}(H) belonging to the purely-imaginary real or quaternionic sets of operators in eq.˜2.8.

We find that for the non-chiral classes our theorem is complete via Section˜4 and in particular the results there which are summarized in Table˜3.

5.5 Classification of the chiral classes

Now we assume that Π\Pi is present, i.e., that we are in such AZ classes where insulators obey {H,Π}=0\{H,\Pi\}=0. Thanks to Assumption˜5.10, it must be that N=2WN=2W for some W1W\in\mathbb{N}_{\geq 1}, and so the Hilbert space breaks into a direct sum

=(2()W)(2()W).\displaystyle\mathcal{H}=\left(\ell^{2}(\mathbb{Z})\otimes\mathbb{C}^{W}\right)\oplus\left(\ell^{2}(\mathbb{Z})\otimes\mathbb{C}^{W}\right)\,.

We formally refer to the left copy as “positive chirality” and the other as “negative chirality”, and use ±\mathcal{H}_{\pm} for these two. Since they are isomorphic in a local way, we will actually drop the distinction between them. By a local (at the level of N\mathbb{C}^{N}) unitary transformation on \mathcal{H} we may without loss of generality assume that Π\Pi is diagonal, i.e., acting as 𝟙W(𝟙W)\mathds{1}_{W}\oplus(-\mathds{1}_{W}) on each local copy of WW\mathbb{C}^{W}\oplus\mathbb{C}^{W}. Hence it must be that insulators which are chiral have the form

H=[0SS0]\displaystyle H=\begin{bmatrix}0&S^{\ast}\\ S&0\end{bmatrix}

for some S(2()W)S\in\mathcal{B}(\ell^{2}(\mathbb{Z})\otimes\mathbb{C}^{W}) which is not necessarily self-adjoint (note how in writing SS we dropped the distinction between the positive and negative chiralities). Moreover, in this chiral grading, Λ\Lambda is diagonal.

Clearly, the spectral gap condition on HH translates to SS being invertible since |H|2=|S|2|S|2|H|^{2}=|S|^{2}\oplus|S^{\ast}|^{2}, and HH is local iff SS is. Moreover, via [GS18, Lemma 2],

(5.4) sgn(H)=[0pol(S)pol(S)0]\displaystyle\operatorname{sgn}(H)=\begin{bmatrix}0&\operatorname{pol}(S)^{\ast}\\ \operatorname{pol}(S)&0\end{bmatrix}

where pol(S)S|S|1\operatorname{pol}(S)\equiv S|S|^{-1} is the polar part of SS, which is in our setting unitary since SS is invertible.

Finally, it is interesting to note that

Lemma 5.14.

If HH\in\mathcal{I} has chiral symmetry then it is a bulk-insulator automatically. Thus, the bulk-insulator constraint is vacuous within the chiral classes.

Proof.

Using the same im(Λ)im(Λ)\operatorname{im}(\Lambda)^{\perp}\oplus\operatorname{im}(\Lambda) decomposition of the Hilbert space as in eq.˜4.4, we write

sgn(H)=[XAAY].\displaystyle\operatorname{sgn}(H)=\begin{bmatrix}X&A\\ A^{\ast}&Y\end{bmatrix}\,.

We caution the reader that only within this proof and unlike the rest of this section, X,YX,Y are blocks in sgn(H)\operatorname{sgn}(H) and not position operators.

Since [Π,Λ]=0[\Pi,\Lambda]=0, we conclude that {X,Π}={Y,Π}=0\Set{X,\Pi}=\Set{Y,\Pi}=0. Thus, both XX and YY have spectra which is symmetric about zero and contained in [1,1][-1,1], and, is discrete on (1,1)(-1,1). But since, by assumption, both im(Λ)\operatorname{im}(\Lambda)^{\perp} and im(Λ)\operatorname{im}(\Lambda) are infinite-dimensional, it cannot be that either XX or YY have only discrete spectrum. As a result, we conclude that both XX and YY have both ±1\pm 1 in their essential spectrum. We conclude by Lemma˜2.3, which implies that X,YX,Y are essentially non-trivial self-adjoint unitaries.

𝔽\mathbb{F} AZ Class Topological invariant
\mathbb{C} AIII indΛpol(S)\operatorname{ind}\mathbb{\Lambda}\operatorname{pol}(S)
\mathbb{R} BDI indΛpol(S)\operatorname{ind}\mathbb{\Lambda}\operatorname{pol}(S)
\mathbb{H} CII indΛpol(S)2\operatorname{ind}\mathbb{\Lambda}\operatorname{pol}(S)\in 2\mathbb{Z}
\star\mathbb{R} CI -
\star\mathbb{H} DIII ind2Λpol(S)\operatorname{ind}_{2}\mathbb{\Lambda}\operatorname{pol}(S)
Table 5: Correspondence between the operators defined in eqs.˜2.6 and 2.7 and the chiral AZ symmetry classes, and formulas for the topological invariants. Here SS stands for the off-diagonal block within HH in the presence of chiral symmetry, and pol\operatorname{pol} is its polar part.
Lemma 5.15.

For Σ\Sigma in the chiral classes, the space B,Σ\mathcal{I}_{B,\Sigma}^{\flat} is homeomorphic to 𝒰𝔽\mathcal{U}_{\mathbb{F}}^{\mathcal{L}} with the correspondence between Σ\Sigma and 𝔽\mathbb{F} as depicted in Table˜5.

Proof.

Most of the necessary statements for the proof have just appeared above so we really only need to focus on the correspondence between the physical symmetry classes of Θ\Theta and Ξ\Xi versus the abstract real and quaternionic operator classes defined in Section˜2.

Clearly for Σ=AIII\Sigma=\mathrm{AIII} the mapping given by

(5.5) B,AIII[0UU0]U𝒰\displaystyle\mathcal{I}_{B,\mathrm{AIII}}^{\flat}\ni\begin{bmatrix}0&U^{\ast}\\ U&0\end{bmatrix}\mapsto U\in\mathcal{U}^{\mathcal{L}}_{\mathbb{C}}

is the required homeomorphism, which is indeed a homeomorphism: well-definedness and bijectivity follow by the foregoing discussion and continuity is clear.

We proceed with the other four choices of Σ\Sigma. By ˜5.9 and Assumption˜5.10, we are left only to check what Θ\Theta squares to and whether it commutes or anti-commutes with Π\Pi: four possibilities. As was explained in ˜5.9, when, in Table˜1, Ξ2\Xi^{2} and Θ2\Theta^{2} disagree we should take {Θ,Π}=0\Set{\Theta,\Pi}=0 and when they agree we take [Θ,Π]=0[\Theta,\Pi]=0.

Let us write the time-reversal symmetry operator in the chiral grading as

Θ=[Θ++Θ+Θ+Θ].\displaystyle\Theta=\begin{bmatrix}\Theta_{++}&\Theta_{+-}\\ \Theta_{-+}&\Theta_{--}\end{bmatrix}\,.
  1. 1.

    When [Π,Θ]=0[\Pi,\Theta]=0 (classes BDI and CII) we have Θ+=Θ+=0\Theta_{+-}=\Theta_{-+}=0. In this case, we define F:=Θ++=ΘF:=\Theta_{++}=\Theta_{--} (they are the same by Assumption˜5.10) and we find that under the mapping eq.˜5.5 the condition [sgn(H),Θ]=0[\operatorname{sgn}(H),\Theta]=0 implies UF=FUUF=FU, i.e., UU is either a real or quaternionic operator based on the value of F2F^{2}: for Θ2=𝟙\Theta^{2}=\mathds{1} (class BDI) we get U𝒰U\in\mathcal{U}^{\mathcal{L}}_{\mathbb{R}} and for Θ2=𝟙\Theta^{2}=-\mathds{1} (class CII) we get U𝒰U\in\mathcal{U}^{\mathcal{L}}_{\mathbb{H}}.

  2. 2.

    When {Π,Θ}=0\Set{\Pi,\Theta}=0 (classes DIII and CI) we have Θ++=Θ=0\Theta_{++}=\Theta_{--}=0. Assumption˜5.10 allows us further to avoid notation overhead since Θ+=Θ+=:F\Theta_{+-}=-\Theta_{-+}^{\ast}=:F. In this case, however, [sgn(H),Θ]=0[\operatorname{sgn}(H),\Theta]=0 implies UF=FUUF=FU^{\ast}, which is precisely the \star-real or \star-quaternionic condition, based on F2=±𝟙F^{2}=\pm\mathds{1}, which is equal to the value of Θ2\Theta^{2}. Hence we find that for Θ2=𝟙\Theta^{2}=\mathds{1} (class CI) U𝒰U\in\mathcal{U}^{\mathcal{L}}_{\star\mathbb{R}} and for Θ2=𝟙\Theta^{2}=-\mathds{1} (class DIII), U𝒰U\in\mathcal{U}^{\mathcal{L}}_{\star\mathbb{H}}.

Now as a result of the statements in Section˜3, the proof of Theorem˜5.12 is complete.

5.6 Classification of exponentially local chiral insulators

Our theory so far has involved the one-dimensional locality condition ˜5.1. This condition may appear somewhat contrived from the physical stand point, in the sense that all it asks is that Hamiltonians HH obey ΛHΛ𝒦\Lambda H\Lambda^{\perp}\in\mathcal{K}. This condition may be criticized (and we would agree, rightly so) that too much of the physics has been washed away.

In this subsection we address this issue as follows: we consider one-dimensional operators with exponential locality as in eq.˜5.1, but only in class AIII for simplicity. Indeed, this type of endeavor is somewhat perpendicular to the activity of topological classification, and is more related to a study of regularity and approximation. In the commutative setting this would be tantamount to a type of Whitney approximation theorem saying that for any two smooth manifolds X,YX,Y with Y=\partial Y=\varnothing, any continuous map XYX\to Y is continuously homotopic to a smooth map XYX\to Y. For that reason we restrict ourselves here merely to one non-trivial symmetry class rather than repeat the analysis for all the AZ classes.

Hence, let us define

Definition 5.16 (exponentially local insulators).

An exponentially local insulator is a self-adjoint Hamiltonian H=HH=H^{\ast}\in\mathcal{B} which is spectrally gapped (at zero) and for which exponential locality eq.˜5.1 holds with any rate:

infx,yd1xylog(Hxy)>0.\displaystyle\inf_{x,y\in\mathbb{Z}^{d}}-\frac{1}{\left\lVert x-y\right\rVert}\log\left(\left\lVert H_{xy}\right\rVert\right)>0\,.

We denote this space by exp\mathcal{I}_{\mathrm{exp}} and furnish it with the subspace topology from the operator norm topology. We note that at least for the chiral classes there is no need to speak of the bulk-insulator condition thanks to Lemma˜5.14 (recall exponentially local operators are Λ\Lambda-local).

It is a fact that this space is strictly smaller than the one obtained with ˜5.1. Indeed, an explicit example may be constructed [Gei22, Example 3.3.10].

The classification result for exponentially local chiral operators is thus:

Theorem 5.17 (AIII d=1d=1 exp. local classification).

In d=1d=1, at fixed NN\in\mathbb{N}, the space exp,AIII\mathcal{I}_{\mathrm{exp,AIII}} has \mathbb{Z} path components labeled by the norm continuous map

exp,AIII[0SS0]indΛS.\displaystyle\mathcal{I}_{\mathrm{exp,AIII}}\ni\begin{bmatrix}0&S^{\ast}\\ S&0\end{bmatrix}\mapsto\operatorname{ind}\mathbb{\Lambda}S\in\mathbb{Z}\,.

We note that in this theorem, it is easier for us to deform within the space of exponentially local invertible operators rather than exponentially local unitary operators as in the preceding proof. This is no major disadvantage though, since our ultimate goal is exp,AIII\mathcal{I}_{\mathrm{exp,AIII}} rather than exp,AIII\mathcal{I}_{\mathrm{exp,AIII}}^{\flat}.

Proof.

Similarly to the proof of Theorem˜3.1, the continuity, surjectivity and logarithmic law for indΛ\operatorname{ind}\circ\mathbb{\Lambda} are established, so we are really only concerned with injectivity of the map at the level of the path-components.

Hence, let S(2()W)S\in\mathcal{B}(\ell^{2}(\mathbb{Z})\otimes\mathbb{C}^{W}) be invertible and have zero index, and our goal is to continuously connect it to 𝟙\mathds{1} within the space of exponentially local invertibles.

Let RR be the bilateral right shift operator on 2()\ell^{2}(\mathbb{Z}). Then clearly we may write

S=lSlRl\displaystyle S=\sum_{l\in\mathbb{Z}}S_{l}R^{l}

where for each ll\in\mathbb{Z}, SlS_{l} is a diagonal operator given via its matrix elements

(Sl)xy=δxySx,xl(x,y).\displaystyle(S_{l})_{xy}=\delta_{xy}S_{x,x-l}\qquad(x,y\in\mathbb{Z})\,.

The series converges in operator norm thanks to exponential locality. Hence, given any ε>0\varepsilon>0, there is some Lε>0L_{\varepsilon}>0 such that the LεL_{\varepsilon} hopping operator SLε:=|l|LεSlRlS^{L_{\varepsilon}}:=\sum_{|l|\leq L_{\varepsilon}}S_{l}R^{l} is ε\varepsilon-close to SS:

SSLε<ε.\displaystyle\left\lVert S-S^{L_{\varepsilon}}\right\rVert<\varepsilon\,.

Moreover, the straight line homotopy

[0,1]t(1t)S+tSLε\displaystyle[0,1]\ni t\mapsto(1-t)S+tS^{L_{\varepsilon}}

is clearly norm continuous, and passes through locals. It passes through invertibles too if ε\varepsilon is chosen sufficiently small since these are open. This shows that without loss of generality we may assume that SS is of finite hopping.

Next, for any finite hopping operator, there is an integer W~\widetilde{W} (in particular, W~=W×Lε\widetilde{W}=W\times L_{\varepsilon}) and a local unitary transformation

U:2()W~2()W\displaystyle U:\ell^{2}(\mathbb{Z})\otimes\mathbb{C}^{\widetilde{W}}\to\ell^{2}(\mathbb{Z})\otimes\mathbb{C}^{W}

which does not affect the Λ\Lambda-index (this is “redimerization”) such that

USLεU=A+BR+CR\displaystyle U^{\ast}S^{L_{\varepsilon}}U=A+BR+CR^{\ast}

where A~,B~,C~\widetilde{A},\widetilde{B},\widetilde{C} are diagonal operators (so, sequences W~\mathbb{Z}\to\mathbb{C}^{\widetilde{W}}). Indeed this map is

U(ψ)(,[ψ0ψLε1],[ψLεψ2Lε1],[ψ2Lεψ3Lε1],).\displaystyle U^{\ast}(\psi)\equiv(\dots,\begin{bmatrix}\psi_{0}\\ \dots\\ \psi_{L_{\varepsilon}-1}\end{bmatrix},\begin{bmatrix}\psi_{L_{\varepsilon}}\\ \dots\\ \psi_{2L_{\varepsilon}-1}\end{bmatrix},\begin{bmatrix}\psi_{2L_{\varepsilon}}\\ \dots\\ \psi_{3L_{\varepsilon}-1}\end{bmatrix},\dots)\,.

Let us factor out a left shift operator

A+BR+CR=(AR+BR2+C)R\displaystyle A+BR+CR^{\ast}=(AR+BR^{2}+C)R^{\ast}

and consider another redimerization V:2()2W~2()W~V:\ell^{2}(\mathbb{Z})\otimes\mathbb{C}^{2\widetilde{W}}\to\ell^{2}(\mathbb{Z})\otimes\mathbb{C}^{\widetilde{W}} so that we can write

V(AR+BR2+C)V=A^+B^R\displaystyle V^{*}(AR+BR^{2}+C)V=\widehat{A}+\widehat{B}R

where A^,B^:2W~\widehat{A},\widehat{B}:\mathbb{Z}\to\mathbb{C}^{2\widetilde{W}} are diagonal operators (the transition from A+BR+CRA+BR+CR^{\ast} to A+BRA+BR was described in [GS18, Example 1]). In particular

indΛ(A^+B^R)=W~\displaystyle\operatorname{ind}\mathbb{\Lambda}(\widehat{A}+\widehat{B}R)=-\widetilde{W}

due to the factoring-out of a left shift operator which has index +W~+\widetilde{W}. Using Lemma˜5.18 below, we can deform A^+B^R\widehat{A}+\widehat{B}R to D+DRD^{\perp}+DR where D:2W~D:\mathbb{Z}\to\mathbb{C}^{2\widetilde{W}} is diagonal with

Dn=[0W~00𝟙W~](n)\displaystyle D_{n}=\begin{bmatrix}0_{\widetilde{W}}&0\\ 0&\mathds{1}_{\widetilde{W}}\end{bmatrix}\qquad(n\in\mathbb{Z})

and the deformation is within the space of invertible operators of the same nearest-neighbor form as A^+B^R\widehat{A}+\widehat{B}R. In conclusion, we have the path

USLεU=A+BR+CRV(D+DR)VR\displaystyle U^{\ast}S^{L_{\varepsilon}}U=A+BR+CR^{\ast}\rightsquigarrow V(D^{\perp}+DR)V^{*}R^{*}

within the space of exponentially local invertibles. This last operator, however, is readily seen to be equal to

[0𝟙W~𝟙W~0][0𝟙W~𝟙W~0].\displaystyle\dots\oplus\begin{bmatrix}0&\mathds{1}_{\widetilde{W}}\\ \mathds{1}_{\widetilde{W}}&0\end{bmatrix}\oplus\begin{bmatrix}0&\mathds{1}_{\widetilde{W}}\\ \mathds{1}_{\widetilde{W}}&0\end{bmatrix}\oplus\dots\,.

Each such 2×22\times 2 block may be deformed (by a local Kuiper) to 𝟙2W~\mathds{1}_{2\widetilde{W}} within the space 𝒰(2W~)\mathcal{U}(2\widetilde{W}) and hence the whole operator to 𝟙\mathds{1} within the infinite 𝒰\mathcal{U} (so that, in particular, this deformation passes within invertibles and exponentially locals). We have thus exhibited a path USLεU𝟙2()W~U^{\ast}S^{L_{\varepsilon}}U\rightsquigarrow\mathds{1}_{\ell^{2}(\mathbb{Z})\otimes\mathbb{C}^{\widetilde{W}}}. Conjugating that path with UU and composing with the straight line path from SS to SLεS^{L_{\varepsilon}} we obtain a path in the original Hilbert space S𝟙2()WS\rightsquigarrow\mathds{1}_{\ell^{2}(\mathbb{Z})\otimes\mathbb{C}^{W}}. ∎

Lemma 5.18 (A+BRA+BR homotopies).

On the space 2()K\ell^{2}(\mathbb{Z})\otimes\mathbb{C}^{K}, let RR be the bilateral right-shift operator on 2()\ell^{2}(\mathbb{Z}) and A,B:MatK()A,B:\mathbb{Z}\to\operatorname{Mat}_{K}(\mathbb{C}) be diagonal operators. For any kk\in\mathbb{Z}, the space of invertible operators on 2()K\ell^{2}(\mathbb{Z})\otimes\mathbb{C}^{K} of the form A+BRA+BR having indΛ(A+BR)=k\operatorname{ind}_{\Lambda}(A+BR)=k is path-connected.

Proof.

We first note that if A+BRA+BR is invertible, then A+λBRA+\lambda BR is invertible for any λ𝕊1\lambda\in\mathbb{S}^{1}. To see this, for any λ𝕊1\lambda\in\mathbb{S}^{1} we define the diagonal operator Uλ:=λX𝟙KU_{\lambda}:=\lambda^{X}\otimes\mathds{1}_{K} where XX is the position operator on 2()\ell^{2}(\mathbb{Z}). I.e.,

(Uλ)nm:=δnmλn𝟙K(n,m).\displaystyle(U_{\lambda})_{nm}:=\delta_{nm}\lambda^{n}\mathds{1}_{K}\qquad(n,m\in\mathbb{Z})\,.

Since RXR=X+𝟙R^{\ast}XR=X+\mathds{1} we have RUλR=λUλR^{\ast}U_{\lambda}R=\lambda U_{\lambda}. Thus, using the fact that the diagonal operators A,BA,B commute with UλU_{\lambda} we find

A+λBR=A+BR(RUλRUλ)=λ𝟙=AUλUλ+BUλRUλ=Uλ(A+BR)Uλ\displaystyle A+\lambda BR=A+BR\underbrace{\left(R^{\ast}U_{\lambda}RU_{\lambda}^{\ast}\right)}_{=\lambda\mathds{1}}=AU_{\lambda}U_{\lambda}^{\ast}+BU_{\lambda}RU_{\lambda}^{\ast}=U_{\lambda}\left(A+BR\right)U_{\lambda}^{\ast}

from which we conclude A+λBRA+\lambda BR is invertible for all λ𝕊1\lambda\in\mathbb{S}^{1}.

It is hence justified to apply Theorem˜C.2 below, with G=BRG=BR to obtain idempotents P,QP,Q with respect to which AA and BRBR are diagonal according to the grading eq.˜C.2, that is, we have

A\displaystyle A =QAP+QAP\displaystyle=Q^{\perp}AP^{\perp}+QAP
BR\displaystyle BR =QBRP+QBRP\displaystyle=Q^{\perp}BRP^{\perp}+QBRP

(with P𝟙PP^{\perp}\equiv\mathds{1}-P, despite not being orthogonal) and

(5.6) λ𝔻¯,{Q(A+λBR)P:imPimQQ(λA+BR)P:imPimQ are both invertible.\displaystyle\forall\lambda\in\overline{\mathbb{D}}\,,\qquad\begin{cases}Q^{\perp}(A+\lambda BR)P^{\perp}&:\operatorname{im}P^{\perp}\to\operatorname{im}Q^{\perp}\\ \quad Q(\lambda A+BR)P&:\operatorname{im}P\to\operatorname{im}Q\end{cases}\text{ are both invertible.}

We therefore consider the path

(5.7) [0,1]tQ(A+(1t)BR)P+Q((1t)A+BR)P\displaystyle[0,1]\ni t\mapsto Q^{\perp}(A+(1-t)BR)P^{\perp}+Q((1-t)A+BR)P

which is invertible by construction (each of its two diagonal blocks is separately invertible), and as we shall see, is also of the A+BRA+BR-hopping form. Indeed, we argue that the idempotents P,QP,Q are diagonal in space. We note in passing that this fact was observed in [BG91], from which we draw inspiration. Let Δ:\mathbb{\Delta}:\mathcal{B}\to\mathcal{B} be the super-operator which projects an operator to its diagonal part, i.e., for any operator FF\in\mathcal{B},

(ΔF)nmFnmδnm(n,m).\displaystyle(\mathbb{\Delta}F)_{nm}\equiv F_{nm}\delta_{nm}\qquad(n,m\in\mathbb{Z})\,.

Clearly R(ΔF)R=Δ(RFR)R^{\ast}\left(\mathbb{\Delta}F\right)R=\mathbb{\Delta}\left(R^{\ast}FR\right). Then we note the Cauchy identity

12πi𝕊1UλFUλdλ=RΔF\displaystyle\frac{1}{2\pi\operatorname{i}}\oint_{\mathbb{S}^{1}}U_{\lambda}FU_{\lambda}^{*}\operatorname{d}{\lambda}=R^{\ast}\mathbb{\Delta}F

which follows from the calculation on the n,mn,m matrix elements

(12πi𝕊1UλFUλdλ)nm=(12πi𝕊1λnmdλ)Fnm=δnm,1Fnm=(RΔF)nm\displaystyle\left(\frac{1}{2\pi\operatorname{i}}\oint_{\mathbb{S}^{1}}U_{\lambda}FU_{\lambda}^{*}\operatorname{d}{\lambda}\right)_{nm}=\left(\frac{1}{2\pi\operatorname{i}}\oint_{\mathbb{S}^{1}}\lambda^{n-m}\operatorname{d}{\lambda}\right)F_{nm}=\delta_{n-m,-1}F_{nm}=\left(R^{\ast}\mathbb{\Delta}F\right)_{nm}

and so using the definition of PP from eq.˜C.1 we have

P12πi𝕊1(A+λBR)1BRdλ=12πi𝕊1Uλ(A+BR)1UλdλBR\displaystyle P\equiv\frac{1}{2\pi\operatorname{i}}\oint_{\mathbb{S}^{1}}(A+\lambda BR)^{-1}BR\operatorname{d}{\lambda}=\frac{1}{2\pi\operatorname{i}}\oint_{\mathbb{S}^{1}}U_{\lambda}(A+BR)^{-1}U_{\lambda}^{\ast}\operatorname{d}{\lambda}BR =R(Δ(A+BR)1)BR\displaystyle=R^{\ast}\left(\mathbb{\Delta}\left(A+BR\right)^{-1}\right)BR
=Δ(R(A+BR)1BR).\displaystyle=\mathbb{\Delta}\left(R^{\ast}\left(A+BR\right)^{-1}BR\right)\,.

Similarly, the idempotent Q=ΔB(A+BR)1Q=\mathbb{\Delta}B\left(A+BR\right)^{-1} and is in particular diagonal. Therefore, eq.˜5.7 indeed passes within the space of operators of the form A+BRA+BR and yields, with the diagonal idempotent Π:=RPR=Δ((A+BR)1B)\Pi:=RPR^{\ast}=\mathbb{\Delta}\left(\left(A+BR\right)^{-1}B\right), the path

A+BRQAP+QBRP=QAP+QBΠR.\displaystyle A+BR\rightsquigarrow Q^{\perp}AP^{\perp}+QBRP=Q^{\perp}AP^{\perp}+QB\Pi R\,.

Our goal now is to further deform the two diagonal operators QAP,QBΠQ^{\perp}AP^{\perp},QB\Pi into D,DD^{\perp},D where

(5.8) D=𝟙M\displaystyle D=\mathds{1}\otimes M

for some constant K×KK\times K matrix given by M:=0Kp𝟙pM:=0_{K-p}\oplus\mathds{1}_{p} for some p=0,,Kp=0,\dots,K.

To get to D+DRD^{\perp}+DR, we note that QBΠ:imΠimQQB\Pi:\operatorname{im}\Pi\to\operatorname{im}Q and QAP:imPimQQ^{\perp}AP^{\perp}:\operatorname{im}P^{\perp}\to\operatorname{im}Q^{\perp} are both invertible since they are the point λ=0\lambda=0 of eq.˜5.6. Since all these operators are diagonal, they define a sequence of invertible maps QnBnPn1:imPn1imQnQ_{n}B_{n}P_{n-1}:\operatorname{im}P_{n-1}\to\operatorname{im}Q_{n} and QnAPn:imPnimQnQ_{n}^{\perp}AP_{n}^{\perp}:\operatorname{im}P_{n}^{\perp}\to\operatorname{im}Q_{n}^{\perp}. For every nn\in\mathbb{Z}, the invertibility of the first matrix implies that dimimPn1=dimQn\dim\operatorname{im}P_{n-1}=\dim Q_{n} and of the second dimimPn=dimimQn\dim\operatorname{im}P_{n}^{\perp}=\dim\operatorname{im}Q_{n}^{\perp} and hence dimimPn=dimimQn\dim\operatorname{im}P_{n}=\dim\operatorname{im}Q_{n} so that

ndimimPn=dimimQn\displaystyle\mathbb{Z}\ni n\mapsto\dim\operatorname{im}P_{n}=\dim\operatorname{im}Q_{n}\in\mathbb{N}

is a constant sequence, whose constant value we denote by pp\in\mathbb{N}. Our goal is to deform the two sequences of invertible maps above into the trivial ones given by nM,Mn\mapsto M,M^{\perp}; to do one has to deform both the vector subspaces on which these maps are invertible so as to make them constant in nn (this amounts to a unitary conjugation of Pn,QnP_{n},Q_{n}), and then deform the maps within the respective subspaces to 𝟙\mathds{1} resp. 0. Indeed, the constant rank of Qn,PnQ_{n},P_{n} implies that we can find unitaries U,VU,V which are diagonal in space and so that P=UDUP=U^{\ast}DU and Q=VDVQ=V^{\ast}DV. As such, for each nn\in\mathbb{Z}, the K×KK\times K matrices (VAU)n(VAU^{\ast})_{n} and (VBRUR)n(VBRU^{\ast}R^{*})_{n} both restrict to invertible maps imMimM\operatorname{im}M^{\perp}\to\operatorname{im}M^{\perp} and imMimM\operatorname{im}M\to\operatorname{im}M respectively. Using Kuiper we deform each of these matrices to MM^{\perp} and MM respectively to obtain a local invertible deformation A+BRD+DRA+BR\rightsquigarrow D^{\perp}+DR.

Let us see that the path we describe above indeed passes through invertibles. We have so far

A+BR\displaystyle A+BR QAP+QBRP\displaystyle\rightsquigarrow Q^{\perp}AP^{\perp}+QBRP
=(VDV)A(UDU)+(VDV)BR(UDU)\displaystyle=(V^{*}D^{\perp}V)A(U^{*}D^{\perp}U)+(V^{*}DV)BR(U^{*}DU)
=V(DVAUD)U+V(DVBRUD)U\displaystyle=V^{*}(D^{\perp}VAU^{*}D^{\perp})U+V^{*}(DVBRU^{*}D)U
D(VAU)D+D(VBRU)D\displaystyle\rightsquigarrow D^{\perp}(VAU^{*})D^{\perp}+D(VBRU^{*})D

where the last deformation follows by a Kuiper to get V,U𝟙V,U\rightsquigarrow\mathds{1}. Now we note

VAU:kerDkerD,VBRU:imDimD\displaystyle VAU^{*}:\ker D\to\ker D,\quad VBRU^{*}:\operatorname{im}D\to\operatorname{im}D

are invertible by construction. We note that VBRUVBRU^{*} is not diagonal and shall be deformed to DRDR rather than DD. To this end, write VBRU=(VBRUR)RVBRU^{*}=(VBRU^{*}R^{*})R and note VBRUR:imDimDVBRU^{*}R^{*}:\operatorname{im}D\to\operatorname{im}D is invertible and diagonal. Deform within diagonal invertibles in each respective subspace

VAUD:kerDkerD,VBRURD:imDimD\displaystyle VAU^{*}\rightsquigarrow D^{\perp}:\ker D\to\ker D,\quad VBRU^{*}R^{*}\rightsquigarrow D:\operatorname{im}D\to\operatorname{im}D

Then (VBRUR)RDR:imDimD(VBRU^{*}R^{*})R\rightsquigarrow DR:\operatorname{im}D\to\operatorname{im}D within invertibles. Thus

D(VAU)D+D(VBRU)D\displaystyle D^{\perp}(VAU^{*})D^{\perp}+D(VBRU^{*})D D(D)D+D(DR)D\displaystyle\rightsquigarrow D^{\perp}(D^{\perp})D^{\perp}+D(DR)D
=D+DRD=D+DR\displaystyle=D^{\perp}+DRD=D^{\perp}+DR

since RD=DRRD=DR. The path is invertible since we treat kerDkerD\ker D\to\ker D and imDimD\operatorname{im}D\to\operatorname{im}D separately.

Since indΛ(A+BR)=k\operatorname{ind}_{\Lambda}(A+BR)=k is preserved under norm continuous deformations, its value is equal to indΛ(D+DR)=pK\operatorname{ind}_{\Lambda}(D^{\perp}+DR)=p-K, where we identified p=K+k.p=K+k. Any two invertible operators S,TS,T of the form A+BRA+BR with the same index can be deformed to the same D+DRD^{\perp}+DR, and hence can be deformed to each other. ∎

6 Classification of one-dimensional edge systems

A prominent feature of topological insulators is the bulk-edge correspondence, which states, roughly speaking, that “the topology” of infinite systems agrees with the topology of the associated systems truncated to the half-space. This vague statement has physical content (about existence of edge modes) and two mathematical assertions: that the topological classifications of these two types of geometries agree, and moreover, that given a bulk insulator HH, if we were to truncate it to the half-space (with largely any reasonable boundary conditions) to get H^\widehat{H}, calculating the index for HH or for H^\widehat{H} (using different formulas) would yield the same number. This latter, numerical as it were, type of bulk-edge correspondence has been the subject of many papers, starting with the integer quantum Hall effect [Hat93], and continuing with the more mathematical [SKR00, EG02]; As far as we are aware, that the two topological classifications agree (without numerical equivalence) has been established for the entire table using KK-theory [BKR17] in the spectral gap regime.

Let us make a few comments about the edge classification in the current setting. The one-dimensional edge Hilbert space is 2()N\ell^{2}(\mathbb{N})\otimes\mathbb{C}^{N}. Now, the constraint of locality which was presented in ˜5.1 does not make sense anymore in the edge (unlike locality in the form of eq.˜5.1 which would carry over directly). Moreover, generically edge systems are not insulators: rather, they are truncations of infinite systems which are bulk insulators. In the spectral gap regime this may be encoded with or without recourse to a bulk Hamiltonian, as presented in [BSS23, Section 2.4]. In the one dimensional spectral gap setting, however, the situation is somewhat simplified for the following reason: by adding a truncation, we may only create finite-degeneracy eigenvalues but not change the essential spectrum (since the truncation is a compact perturbation of the bulk system). However, according to the RAGE theorem, eigenvalues are exponentially decaying from some center, and thus exponentially decaying from the truncation. As a result, it would appear that asking that the edge Hamiltonian is a Fredholm operator suffices for the bulk-gap requirement, because Fredholm operators are precisely those which are essentially gapped at zero.

But more is true: the Fredholm condition is a very weak notion of locality which in the edge setting is a good replacement for ˜5.1. Indeed, if we think of the Fredholm condition as the finiteness of the kernel and the kernel of the adjoint, this is essentially asking that the operator cannot have too far away hopping, since if it did, that would violate the finite kernel condition.

As such, it would appear that in one dimension, locality and the gap condition collapse into one insulator condition:

Definition 6.1 (one-dimensional edge insulators).

H^=H^(2()N)\widehat{H}=\widehat{H}^{\ast}\in\mathcal{B}(\ell^{2}(\mathbb{N})\otimes\mathbb{C}^{N}) is an edge insulator (with bulk gap at zero energy) iff it is a Fredholm operator.

Clearly, in the edge picture we do not need to worry about the “bulk” insulator condition but we do need to make sure our systems are non-trivial in the sense that they have essential spectrum below and above zero. This corresponds to Atiyah and Singer’s notion of the non-trivial component \mathcal{F}_{\star}. We conclude that in the one-dimensional edge picture, if we are willing to accept a very weak notion of locality (but we emphasize it has not been completely ignored) the theory reduces to the classical Atiyah-Singer classification of Fredholm operators with symmetries [AS69]. Then, for example, class A corresponds to the non-trivial self-adjoint Fredholm operators sa\mathcal{F}^{\mathrm{sa}}_{\star} and from [AS69] we have

π0(sa)[{0}sa]K1({0}){0}\displaystyle\pi_{0}(\mathcal{F}^{\mathrm{sa}}_{\star})\cong\left[\Set{0}\to\mathcal{F}^{\mathrm{sa}}_{\star}\right]\cong K_{1}(\Set{0})\cong\Set{0}

whereas in class AIII, the chiral off-diagonal sub-block SS must be Fredholm, which automatically implies that [0SS0]\begin{bmatrix}0&S^{\ast}\\ S&0\end{bmatrix} is in sa\mathcal{F}^{\mathrm{sa}}_{\star}. This then reduces to the even older Atiyah-Jänich theorem:

π0()[{0}]K0({0}).\displaystyle\pi_{0}(\mathcal{F})\cong\left[\Set{0}\to\mathcal{F}\right]\cong K_{0}(\Set{0})\cong\mathbb{Z}\,.

One could then phrase an edge analog of Theorem˜5.12; the formulas for the edge indices are obvious: they are the Fredholm indices or the 2\mathbb{Z}_{2} Atiyah-Singer indices of the various Fredholm operators without taking sgn\operatorname{sgn} or polar part and without the application of Λ\mathbb{\Lambda}, according to Tables˜5 and 4. We find:

Theorem 6.2.

One dimensional edge insulators as in ˜6.1 have path components given by the d=1d=1 column of Table˜1, and hence the bulk and edge one dimensional systems have the same classifications. For any given bulk insulator HH, the bulk index calculated from HH agrees with the edge index calculated from H^\widehat{H} where H^\widehat{H} is any edge insulator obtained by truncating HH to the half-space such that H^\widehat{H} is Fredholm and respects the symmetry constraint. Hence we obtain a numerical bulk-edge correspondence.

Sketch of proof.

As explained in the foregoing paragraphs, the classification result is covered by [AS69]. The numerical bulk-edge correspondence proof, at the spectral-gap level, is covered by the proof provided in [GS18, Section 3]. ∎

7 The mobility gap regime

As mentioned above, a more general mathematical criterion to guarantee zero electric conductance (and thus the insulator condition) is through quantum dynamics rather than via a spectral constraint. Drawing on Anderson localization, in [EGS05] a deterministic condition was formulated for one operator; we quote the equivalent condition given in [BSS23, Definition 2.5]: Let B1(Δ)B_{1}(\Delta) be the space of measurable functions f:f:\mathbb{R}\to\mathbb{C} which are non-constant only within Δ\Delta and are bounded by 11.

Definition 7.1 (mobility gap).

A material H=HH=H^{\ast}\in\mathcal{L} is mobility gapped at zero energy iff there exists some open interval Δ0\Delta\ni 0 such that

  1. 1.

    There exists some μ>0\mu>0 such that for any ε>0\varepsilon>0 there exists some Cε<C_{\varepsilon}<\infty such that

    (7.1) supfB1(Δ)f(H)xyCexp(μxy+εx)(x,yd).\displaystyle\sup_{f\in B_{1}(\Delta)}\left\lVert f(H)_{xy}\right\rVert\leq C\exp(-\mu\left\lVert x-y\right\rVert+\varepsilon\left\lVert x\right\rVert)\qquad(x,y\in\mathbb{Z}^{d})\,.

    Hence f(H)f(H) has exponentially decaying off-diagonal matrix elements whose rate of decay is however not uniform in the diagonal direction. Moreover, this statement is uniform in ff.

  2. 2.

    All eigenvalues of HH within Δ\Delta are uniformly finitely degenerate (the above condition implies σ(H)Δ=σpp(H)Δ\sigma(H)\cap\Delta=\sigma_{\mathrm{pp}}(H)\cap\Delta via the RAGE theorem).

The type of decay condition appearing in eq.˜7.1 has been called weakly-local in [ST19, BSS23]. In one dimension it seems however that ˜5.1 is still weaker.

Furthermore, it is well-known from the theory of Anderson localization (see [SW86] e.g.) that any fixed deterministic energy value is almost-surely not an eigenvalue of an Anderson localized random operator. Hence, in particular, even though in the mobility gap regime there is no reason to assume a spectral gap, or no accumulation of spectrum near zero, it wouldn’t seem unreasonable to assume that zero is not an eigenvalue of HH.

Hence, if instead of taking the stronger eq.˜7.1 we merely setup the mobility gap condition as the minimal dynamical constraint to guarantee the existence of the index, we could come up with the following deterministic condition, which relies still on ˜5.1:

Definition 7.2 (tentative definition for mobility gap in d=1d=1).

A material H=HH=H^{\ast}\in\mathcal{L} is mobility gapped at zero iff zero is not an eigenvalue of HH and if sgn(H)\operatorname{sgn}(H)\in\mathcal{L}. We denote this space by mg,v1\mathcal{I}_{\mathrm{mg,v1}}. Its topology remains to be defined.

This condition (up to strengthening the mode of locality) was the one given in [GS18, Assumptions 1 and 2]. It is clear that such operators still have well-defined indices: the fact zero is not an eigenvalue of HH means that sgn(H)\operatorname{sgn}(H) is actually unitary and not merely a partial isometry. But more is true: the entire proof of Theorem˜5.12 goes through if we skip the step connecting Hamiltonians with flat Hamiltonians! Indeed, all that is required is that operators be unitary or self-adjoint projections.

To connect Hamiltonians and flat Hamiltonians, we might employ the following “abstract nonsense” definition and argument:

We shall make use of two different topologies on mg,v1\mathcal{I}_{\mathrm{mg,v1}}. First, let 𝒯ι:mg,v1\mathcal{T}_{\iota:\mathcal{I}_{\mathrm{mg,v1}}\to\mathcal{B}} be the initial topology on mg,v1\mathcal{I}_{\mathrm{mg,v1}} generated by the mapping ι:mg,v1\iota:\mathcal{I}_{\mathrm{mg,v1}}\to\mathcal{B}, inherited from the operator norm topology on \mathcal{B}; this is by definition the subspace topology. Next, the functional calculus implies there is a map on operators sgn:mg,v1\operatorname{sgn}:\mathcal{I}_{\mathrm{mg,v1}}\to\mathcal{B} which maps Hsgn(H)H\mapsto\operatorname{sgn}(H). Let 𝒯sgn:mg,v1\mathcal{T}_{\operatorname{sgn}:\mathcal{I}_{\mathrm{mg,v1}}\to\mathcal{B}} then be the initial topology on mg,v1\mathcal{I}_{\mathrm{mg,v1}} which is generated by sgn:mg,v1\operatorname{sgn}:\mathcal{I}_{\mathrm{mg,v1}}\to\mathcal{B}, where \mathcal{B} is understood with the operator norm topology.

Importantly, any path mg,v1\mathcal{I}_{\mathrm{mg,v1}} continuous w.r.t. 𝒯sgn:mg,v1\mathcal{T}_{\operatorname{sgn}:\mathcal{I}_{\mathrm{mg,v1}}\to\mathcal{B}} preserves an index, should one exist, as an index is always a function of the flat Hamiltonian sgn(H)\operatorname{sgn}(H): it is either ind2,Λsgn(H)\operatorname{ind}_{2,\Lambda}\operatorname{sgn}(H) in the non-chiral case or, in the chiral case, it is ind(2),ΛP+sgn(H)P|imPimP+\operatorname{ind}_{(2),\Lambda}\left.P_{+}\operatorname{sgn}(H)P_{-}\right|_{\operatorname{im}P_{-}\to\operatorname{im}P_{+}} where P±P_{\pm} are the SA projections onto the ±1\pm 1 eigenspaces of Π\Pi.

Lemma 7.3.

The space mg,v1\mathcal{I}^{\flat}_{\mathrm{mg,v1}} with the topology 𝒯ι:mg,v1mg,v1\mathcal{T}_{\iota:\mathcal{I}^{\flat}_{\mathrm{mg,v1}}\to\mathcal{I}_{\mathrm{mg,v1}}} is a strong deformation retract of the space mg,v1\mathcal{I}_{\mathrm{mg,v1}} taken with the topology 𝒯sgn:mg,v1\mathcal{T}_{\operatorname{sgn}:\mathcal{I}_{\mathrm{mg,v1}}\to\mathcal{B}}.

This statement remains true if we add the bulk-insulator constraint as well as any of the ten AZ symmetry constraints: mg,v1,B,Σ\mathcal{I}_{\mathrm{mg,v1},B,\Sigma}^{\flat} is a strong deformation retraction of mg,v1,B,Σ\mathcal{I}_{\mathrm{mg,v1},B,\Sigma} for any ΣAZ\Sigma\in\mathrm{AZ}. However, in order to avoid notational overhead we stick with the notation mg,v1\mathcal{I}_{\mathrm{mg,v1}} where it is understood that we also take the bulk-insulator condition into the definition, as well as the appropriate symmetry.

Proof.

Define F:mg,v1×[0,1]mg,v1F:\mathcal{I}_{\mathrm{mg,v1}}\times[0,1]\to\mathcal{I}_{\mathrm{mg,v1}} via

F(H,t):=(1t)H+tsgn(H)(Hmg,v1,t[0,1]).\displaystyle F(H,t):=(1-t)H+t\operatorname{sgn}(H)\qquad(H\in\mathcal{I}_{\mathrm{mg,v1}},t\in[0,1])\,.

We only show FF is continuous with respect to 𝒯sgn:mg,v1\mathcal{T}_{\operatorname{sgn}:\mathcal{I}_{\mathrm{mg,v1}}\to\mathcal{B}} (times the Euclidean topology on [0,1][0,1]), as the other two properties have already been shown in Lemma˜5.13. The initial topology is generated by a sub-basis of inverse images of open sets on the co-domain. Hence, to show continuity, it suffices to start with such an open subset, and so let UU be open in \mathcal{B}. Then we seek to show that F1(sgn1(U))F^{-1}(\operatorname{sgn}^{-1}(U)) is open. To that end, using the fact that for all t[0,1]t\in[0,1], sgn(F(H,t))=sgn(H)\operatorname{sgn}(F(H,t))=\operatorname{sgn}(H), we have

F1(sgn1(U))\displaystyle F^{-1}(\operatorname{sgn}^{-1}(U)) ={(H,t)|sgn(F(H,t))U}\displaystyle=\Set{(H,t)}{\operatorname{sgn}(F(H,t))\in U}
={(H,t)|sgn(H)U}\displaystyle=\Set{(H,t)}{\operatorname{sgn}(H)\in U}
=sgn1(U)×[0,1].\displaystyle=\operatorname{sgn}^{-1}(U)\times[0,1]\,.

The set on the last line is manifestly open. ∎

Corollary 7.4 (The one-dimensional column of the Kitaev table w.r.t. v1 of the mobility gap topology).

At each fixed NN, for any ΣAZ\Sigma\in\mathrm{AZ}, the path-connected components of mg,v1,N,B,Σ\mathcal{I}_{\mathrm{mg,v1},N,B,\Sigma} considered with the initial topology 𝒯sgn:mg,v1\mathcal{T}_{\operatorname{sgn}:\mathcal{I}_{\mathrm{mg,v1}}\to\mathcal{B}}, agree with the set appearing in the first column of Table˜1.

Proof.

Having the deformation retract, we know that π0(mg,v1)\pi_{0}(\mathcal{I}^{\flat}_{\mathrm{mg,v1}}) is the same as π0(mg,v1)\pi_{0}(\mathcal{I}_{\mathrm{mg,v1}}) where we take the subspace topology for the former and 𝒯sgn:mg,v1\mathcal{T}_{\operatorname{sgn}:\mathcal{I}_{\mathrm{mg,v1}}\to\mathcal{B}} for the latter. That subspace topology 𝒯ι:mg,v1mg,v1\mathcal{T}_{\iota:\mathcal{I}^{\flat}_{\mathrm{mg,v1}}\to\mathcal{I}_{\mathrm{mg,v1}}} equals the topology 𝒯sgn:mg,v1\mathcal{T}_{\operatorname{sgn}:\mathcal{I}^{\flat}_{\mathrm{mg,v1}}\to\mathcal{B}} on mg,v1\mathcal{I}^{\flat}_{\mathrm{mg,v1}}. Indeed, this follows by the transitive property of the initial topology [Gro73, p. 2]: if SXS\subseteq X and f:XYf:X\to Y then the subspace topology on SS from XX taken with the initial topology generated by ff equals the initial topology on SS generated by fιf\circ\iota where ι:SX\iota:S\to X is the inclusion map. But now, 𝒯sgn:mg,v1\mathcal{T}_{\operatorname{sgn}:\mathcal{I}^{\flat}_{\mathrm{mg,v1}}\to\mathcal{B}} and 𝒯ι:mg,v1\mathcal{T}_{\iota:\mathcal{I}^{\flat}_{\mathrm{mg,v1}}\to\mathcal{B}} coincide. This is a consequence of the fact that sgn:mg,v1\operatorname{sgn}:\mathcal{I}_{\mathrm{mg,v1}}^{\flat}\to\mathcal{B} reduces to the inclusion map ι:mg,v1\iota:\mathcal{I}^{\flat}_{\mathrm{mg,v1}}\to\mathcal{B} (since sgnsgn=sgn\operatorname{sgn}\circ\operatorname{sgn}=\operatorname{sgn}), and the subspace topology is precisely generated by the inclusion map.

The end conclusion is that we may calculate π0(mg,v1)\pi_{0}(\mathcal{I}^{\flat}_{\mathrm{mg,v1}}) w.r.t. the topology 𝒯ι:mg,v1\mathcal{T}_{\iota:\mathcal{I}^{\flat}_{\mathrm{mg,v1}}\to\mathcal{B}}, i.e. π0()\pi_{0}(\mathcal{I}^{\flat}) with the subspace topology from norm topology. This, however, is precisely the calculation already done in the proof of Theorem˜5.12. ∎

With this, it would appear that the mobility gap problem is solved in one dimension. We maintain this is, however, not the case. Indeed, a subtlety appears from the fact we allowed ourselves to shift Hamiltonians to always place the Fermi energy at zero, which has thus made the above analysis single out zero energy. This is of course invalid because if we were to ask that all given fixed energies are almost-surely not an eigenvalue we would constrain our operators to have a spectral gap, which we are precisely trying to avoid. So by always placing μ=0\mu=0 we are not allowed to ask that zero is not an eigenvalue, and so, following the theory of Anderson localization, we would make another attempt as

Definition 7.5 (another tentative definition for mobility gap in d=1d=1).

A material H=HH=H^{\ast}\in\mathcal{L} is mobility gapped at zero iff kerH\ker H is finite dimensional and sgn(H)\operatorname{sgn}(H)\in\mathcal{L}. We denote the space of all such operators as mg,v2\mathcal{I}_{\mathrm{mg,v2}}; its topology remains to be defined.

We note that in this case, sgn(H)\operatorname{sgn}(H) is merely a partial isometry with finite kernel (and so it is Fredholm, sgn(H)\operatorname{sgn}(H) having closed range) and that still Λsgn(H)\mathbb{\Lambda}\operatorname{sgn}(H) is Fredholm.

Unfortunately, with the initial topology 𝒯sgn:mg,v2\mathcal{T}_{\operatorname{sgn}:\mathcal{I}_{\mathrm{mg,v2}}\to\mathcal{B}} from norm topology on \mathcal{B}, this definition is still not good enough, as the following counterexample demonstrates a deviation from the Kitaev table. As a result, to rectify this situation, one should either define another topology on mg,v2\mathcal{I}_{\mathrm{mg,v2}} (e.g. the subspace topology) or start with another space entirely. The first step should be to require an open interval around zero to have finite degeneracy (uniformly). Another possibility would be to place a dynamical constraint. We postpone such investigations to future work.

Example 7.6.

There exist two operators R,Smg,v2,AIIIR,S\in\mathcal{I}_{\mathrm{mg,v2},\mathrm{AIII}} which have the same Λ\Lambda-index and yet there is no continuous path connecting them in the topology 𝒯sgn:mg,v2\mathcal{T}_{\operatorname{sgn}:\mathcal{I}_{\mathrm{mg,v2}}\to\mathcal{B}}.

Proof.

Define RR to be the unitary right shift operator on 2()\ell^{2}(\mathbb{Z}) and Λ\Lambda as above projects to span({δn}n1)\operatorname{span}(\Set{\delta_{n}}_{n\geq 1}). With this, we choose S:=ΛRΛ+ΛRΛS:=\Lambda^{\perp}R\Lambda^{\perp}+\Lambda R\Lambda. As such, it is clear that indΛS=indΛR=1\operatorname{ind}_{\Lambda}S=\operatorname{ind}_{\Lambda}R=-1 since ΛR=ΛS\mathbb{\Lambda}R=\mathbb{\Lambda}S. Moreover, it is clear that both have finite dimensional kernels and co-kernels, and pol(R)=R\operatorname{pol}(R)=R and pol(S)=S\operatorname{pol}(S)=S which are both Λ\Lambda-local. This means that both RR and SS are in mg,v2,AIII\mathcal{I}_{\mathrm{mg,v2},\mathrm{AIII}}.

However, we maintain that no path continuous in 𝒯sgn:mg,v2\mathcal{T}_{\operatorname{sgn}:\mathcal{I}_{\mathrm{mg,v2}}\to\mathcal{B}} can exist in between them. Indeed, one may follow Lemma˜7.3 to show that the space mg,v2\mathcal{I}^{\flat}_{\mathrm{mg,v2}} with the topology 𝒯ι:mg,v2mg,v2\mathcal{T}_{\iota:\mathcal{I}^{\flat}_{\mathrm{mg,v2}}\to\mathcal{I}_{\mathrm{mg,v2}}} is a strong deformation retract of the space mg,v2\mathcal{I}_{\mathrm{mg,v2}} taken with the topology 𝒯sgn:mg,v2\mathcal{T}_{\operatorname{sgn}:\mathcal{I}_{\mathrm{mg,v2}}\to\mathcal{B}}. Indeed, all that is used there is the fact that sgnsgn=sgn\operatorname{sgn}\circ\operatorname{sgn}=\operatorname{sgn} and not the unitarity of the operators. However, now we could argue that, just as we did in the proof of ˜7.4, that for the space mg,v2\mathcal{I}^{\flat}_{\mathrm{mg,v2}}, the topologies 𝒯ι:mg,v2mg,v2\mathcal{T}_{\iota:\mathcal{I}^{\flat}_{\mathrm{mg,v2}}\to\mathcal{I}_{\mathrm{mg,v2}}}, 𝒯sgn:mg,v2\mathcal{T}_{\operatorname{sgn}:\mathcal{I}^{\flat}_{\mathrm{mg,v2}}\to\mathcal{B}} and 𝒯ι:mg,v2\mathcal{T}_{\iota:\mathcal{I}^{\flat}_{\mathrm{mg,v2}}\to\mathcal{B}} coincide. As such, if a path from RR to SS continuous in 𝒯sgn:mg,v2\mathcal{T}_{\operatorname{sgn}:\mathcal{I}_{\mathrm{mg,v2}}\to\mathcal{B}} were to exist, flattening that path, would imply the existence of a path from pol(R)=R\operatorname{pol}(R)=R to pol(S)=S\operatorname{pol}(S)=S within mg,v2\mathcal{I}^{\flat}_{\mathrm{mg,v2}} continuous w.r.t. the topology 𝒯ι:mg,v2\mathcal{T}_{\iota:\mathcal{I}^{\flat}_{\mathrm{mg,v2}}\to\mathcal{B}}. Indeed, this uses the fact that sgn:mg,v2mg,v2\operatorname{sgn}:\mathcal{I}_{\mathrm{mg,v2}}\to\mathcal{I}_{\mathrm{mg,v2}}^{\flat} is continuous with respect to the initial topology. This last path, continuous w.r.t. 𝒯ι:mg,v2\mathcal{T}_{\iota:\mathcal{I}^{\flat}_{\mathrm{mg,v2}}\to\mathcal{B}}, cannot exist.

Indeed, the topology 𝒯ι:mg,v2\mathcal{T}_{\iota:\mathcal{I}^{\flat}_{\mathrm{mg,v2}}\to\mathcal{B}} is just the subspace topology from operator norm topology, on the space of partial isometries with finite kernel and co-kernel, i.e., Fredholm partial isometries. For us, RR has no kernel and no co-kernel, with Fredholm index zero, and SS has a kernel and co-kernel of dimension +1+1, again with Fredholm index zero. However, since we are working in the space of partial isometries, it is well-known [Hal82, Problem 130] that the dimensions of the kernel and of the co-kernel are continuous in the operator norm (rather than merely lower semi-continuous as in the case of Fredholm operators). As such, having a continuous path between these two operators, would violate the continuity of the dimension of the kernels for partial isometric Fredholms. ∎

8 Classification of bulk spectrally-gapped insulators in odd d>1d>1

Our analysis so far has focused on one-dimensional structures. Let us now turn our attention to higher dimensions. We seek an analogous notion of locality as presented in ˜5.1 which would apply in higher dimensions. In their textbook, Prodan and Schulz-Baldes [PS16, Chapter 6] present a construction which they ascribe to [Con94, GVF00] of locality in all dimensions which proceeds as follows.

Let us define k:=d/2k:=d/2 in even dimensions and k:=(d1)/2k:=(d-1)/2 in odd dimensions. In the spirit of ˜5.3, let us (without loss) assume that NN is divisible by 2k2^{k}, so that it actually carries a representation of a Clifford algebra with generators γ1,,γd\gamma_{1},\dots,\gamma_{d} (now considered as N×NN\times N matrices). The Dirac operator is then defined as

(8.1) D:=i=1dXiγi\displaystyle D:=\sum_{i=1}^{d}X_{i}\otimes\gamma_{i}

and now, in higher dimensions, we choose the locality projection Λ\Lambda to be

(8.2) Λ:=12(𝟙+sgn(D)).\displaystyle\Lambda:=\frac{1}{2}\left(\mathds{1}+\operatorname{sgn}(D)\right)\,.

This operator no longer acts trivially in the internal space N\mathbb{C}^{N} factor as was the case in d=1d=1. Hence DD and both Λ\Lambda intertwine space and the internal degrees of freedom in a non-trivial way; we note that if d=1d=1 we get back our choice made in ˜5.1. It should be remarked that in our notation sgn(D)\operatorname{sgn}(D) is a partial isometry which may be extended to a unitary in an obvious way.

Definition 8.1 (locality in higher odd dimensions).

We define an operator A(2(d)N)A\in\mathcal{B}(\ell^{2}(\mathbb{Z}^{d})\otimes\mathbb{C}^{N}) (with NN divisible by 2k2^{k} without loss as above and dd odd) to be local iff it is Λ\Lambda-local as in ˜2.5 with the particular choice of Λ\Lambda made in eq.˜8.2.

Going back to ˜5.2, we identify in d=2d=2

sgn(D)=[0exp(iΦ)exp(iΦ)0]\displaystyle\operatorname{sgn}(D)=\begin{bmatrix}0&\exp\left(-\operatorname{i}\Phi\right)\\ \exp\left(\operatorname{i}\Phi\right)&0\end{bmatrix}

with Φarg(X1+iX2)\Phi\equiv\arg(X_{1}+\operatorname{i}X_{2}) the angle-position operator. This d=2d=2 pattern is typical: in even dimensions sgn(D)\operatorname{sgn}(D) breaks into off-diagonal form as above [PS16, Chapter 6]. It is clear that ignoring this internal structure of sgn(D)\operatorname{sgn}(D) we get trivial classification for projections in even dimensions in contradiction to expectations. Hence it is clear that in even dimensions one has to contend with a different notion of locality, one which entails operators which essentially commute with a fixed unitary (the Dirac phase) rather than the Dirac projection. This leads to rather different classification scheme which we have little to say about.

On the other hand, in higher odd dimensions we may proceed by adopting the definition of a bulk insulator as in ˜5.7, i.e., bulk insulators are operators H=HH=H^{\ast}\in\mathcal{L} which have a spectral gap about zero and for which the Fermi projection PP is not merely local but also Λ\Lambda-non-trivial. We emphasize that now, however, it can no longer be reasonably argued that this Λ\Lambda-non-trivial requirement would correspond to bulk systems, since now im(Λ)\operatorname{im}(\Lambda) cannot be identified geometrically with an edge system. Be that as it may, one may carry on and in fact obtain all odd-dimensional columns of the Kitaev table in this way, in precisely the same manner as we did in Section˜5.

We thus phrase, without proof, the following

Theorem 8.2.

At each fixed NN, for d2+1d\in 2\mathbb{N}+1, for any ΣAZ\Sigma\in\mathrm{AZ}, the path-connected components of N,B,Σ\mathcal{I}_{N,B,\Sigma} considered with the subspace topology associated with the operator norm topology, agree with the corresponding set appearing in the odd-dimensional columns of Table˜1.

That is, now bulk insulators are defined as in the foregoing paragraph, using the particular choice of compact-commutator locality and bulk-insulator with the choice of Λ\Lambda as in eq.˜8.2.

To prove this theorem, given Section˜5 and the above paragraph, the missing part is explaining how the dimensions cause the symmetry classes to shift which is a shadow of the K-theoretic identity Ki(S2𝒜)=Ki+2(𝒜)K_{i}(S^{2}\mathcal{A})=K_{i+2}(\mathcal{A}) where SS is the suspension of a C-star algebra. The shift does not mix the chiral and non-chiral classes, and furthermore, classes A and AIII are fixed by the shift. So for either the chiral or non-chiral classes, there is a four orbit shuffle that happens as dd+2d\mapsto d+2. To explain this shift one has to allow Θ\Theta and Ξ\Xi to act non-trivially on the Clifford space.

We avoid doing so here because ultimately, we feel that the notion of locality and bulk-insulator derived from this choice of Λ\Lambda is physically contrived. Yes, one could take the point of view that locality and the gap condition may be any sufficiently strong criterion so that the indices are well-defined. But whereas in one-dimension this still made sense with respect to physical real space, in higher-dimensions, we simply cannot find a way to justify this particular choice of Λ\Lambda locality and Λ\Lambda non-triviality. We thus postpone the higher dimensional problem to future work.


Acknowledgments. We are indebted to Gian Michele Graf and Michael Aizenman for stimulating discussions.

Appendix A The Atiyah-Singer 2\mathbb{Z}_{2} index theory

The material in this section was first presented by Atiyah and Singer in [AS69]. Different proofs appeared in [Sch15, Fon+20] but for the sake of completeness we include a short presentation of the theory here, also since the context is somewhat more abstract than the Θ\Theta-odd analysis which was presented in the appendix [Fon+20].

Lemma A.1 (An explicit Diudonné).

Let T()T\in\mathcal{F}\left(\mathcal{H}\right). Then if SBG1(T)S\in B_{\left\lVert G\right\rVert^{-1}}\left(T\right) where GG is any parametrix of TT then S()S\in\mathcal{F}\left(\mathcal{H}\right) too, and

dim(ker(S))=dim(ker(T))dim(im(Z))\displaystyle\dim\left(\ker\left(S\right)\right)=\dim\left(\ker\left(T\right)\right)-\dim\left(\operatorname{im}\left(Z\right)\right)

where Z:ker(T)im(T)Z:\ker\left(T\right)\to\operatorname{im}\left(T\right)^{\perp} is the Schur-complement of SS in the TT-decomposition, i.e.,

Z:=SCASCB(SDB)1SDA\displaystyle Z:=S_{CA}-S_{CB}\left(S_{DB}\right)^{-1}S_{DA}

with A:=ker(T)A:=\ker\left(T\right), B:=ker(T)B:=\ker\left(T\right)^{\perp}, C:=im(T)C:=\operatorname{im}\left(T\right)^{\perp}, D:=im(T)D:=\operatorname{im}\left(T\right).

Proof.

Decomposing =AB=CD\mathcal{H}=A\oplus B=C\oplus D we find T:ABCDT:A\oplus B\to C\oplus D is written in block-operator form as

T=[000TDB]\displaystyle T=\begin{bmatrix}0&0\\ 0&T_{DB}\end{bmatrix}

with TDB:BDT_{DB}:B\to D a vector space isomorphism, and we may also decompose SS as S:ABCDS:A\oplus B\to C\oplus D in block operator form to get

S=[SCASCBSDASDB].\displaystyle S=\begin{bmatrix}S_{CA}&S_{CB}\\ S_{DA}&S_{DB}\end{bmatrix}\,.

Now if ST\left\lVert S-T\right\rVert is sufficiently small then SDBTDB\left\lVert S_{DB}-T_{DB}\right\rVert is sufficiently small so that SDBS_{DB} is also invertible (this may be verified to be true with the upper bound G1\left\lVert G\right\rVert^{-1}), which guarantees that ZZ exists.

Using an LDU decomposition we may write

S=J1(ZSDB)J2\displaystyle S=J_{1}\left(Z\oplus S_{DB}\right)J_{2}

where J1,J2J_{1},J_{2} are two invertible operators, and as such

dim(ker(S))=dim(ker(Z))+dim(ker(SDB))=0.\displaystyle\dim\left(\ker\left(S\right)\right)=\dim\left(\ker\left(Z\right)\right)+\underbrace{\dim\left(\ker\left(S_{DB}\right)\right)}_{=0}\,.

Now apply rank-nullity on Z:ACZ:A\to C to get

dim(ker(Z))+dim(im(Z))=dim(A)=dim(ker(T))\displaystyle\dim\left(\ker\left(Z\right)\right)+\dim\left(\operatorname{im}\left(Z\right)\right)=\dim\left(A\right)=\dim\left(\ker\left(T\right)\right)

which yields the result. ∎

Theorem A.2 (Atiyah-Singer 2\mathbb{Z}_{2} index).

If F()F\in\mathcal{F}_{\star\mathbb{H}}\left(\mathcal{H}\right) (in the sense of Section˜2), i.e., J:J:\mathcal{H}\to\mathcal{H} is an anti-unitary that squares to 𝟙-\mathds{1} and we have

FJ=JF\displaystyle FJ=JF^{\ast}

then

ind2(F)dim(ker(F))mod22\displaystyle\operatorname{ind}_{2}\left(F\right)\equiv\dim\left(\ker\left(F\right)\right)\mod 2\in\mathbb{Z}_{2}

is well-defined, in the sense that if G()G\in\mathcal{F}_{\star\mathbb{H}}\left(\mathcal{H}\right) and FG\left\lVert F-G\right\rVert is sufficiently small, then

ind2(G)=ind2(F).\displaystyle\operatorname{ind}_{2}\left(G\right)=\operatorname{ind}_{2}\left(F\right)\,.
Proof.

Using the same definitions as in the proof above, we have

dim(ker(G))=dim(ker(F))dim(im(Z))\displaystyle\dim\left(\ker\left(G\right)\right)=\dim\left(\ker\left(F\right)\right)-\dim\left(\operatorname{im}\left(Z\right)\right)

with Z:ker(F)im(F)Z:\ker\left(F\right)\to\operatorname{im}\left(F\right)^{\perp} the Schur complement, given by

Z=GCAGCB(GDB)1GDA.\displaystyle Z=G_{CA}-G_{CB}\left(G_{DB}\right)^{-1}G_{DA}\,.

Since FF is \star-quaternionic with respect to JJ, then

JkerF=(imF),J(kerF)=imF.\displaystyle J\ker F=(\operatorname{im}F)^{\perp},\quad J(\ker F)^{\perp}=\operatorname{im}F\,.

In particular, the expressions

GCAJ=JGCA,GDBJ=JGDB,GCBJ=JGDA,GDAJ=JGCB\displaystyle G_{CA}J=JG_{CA}^{*},\quad G_{DB}J=JG_{DB}^{*},\quad G_{CB}J=JG_{DA}^{*},\quad G_{DA}J=JG_{CB}^{*}

make sense and follow directly from GG being \star-quaternionic. It follows that

ZJ=JZ\displaystyle ZJ=JZ^{*}

We now argue that imZ\operatorname{im}Z is even-dimensional. Let us view Z:(kerZ)imZZ:(\ker Z)^{\perp}\to\operatorname{im}Z as an invertible operator that is \star-quaternionic with respect to JJ. Since ZZ:imZimZZZ^{*}:\operatorname{im}Z\to\operatorname{im}Z is self-adjoint, the space imZ\operatorname{im}Z decomposes into eigen-subspaces from ZZZZ^{*}. Let EE be one of the eigen-subspace and take φE\varphi\in E and write ZZφ=λφZZ^{*}\varphi=\lambda\varphi. Clearly φ~0\widetilde{\varphi}\neq 0 since ZZ and JJ are both linear invertible. Let φ~:=ZJφimZ\widetilde{\varphi}:=ZJ\varphi\in\operatorname{im}Z. We have

φ,ZJφ=Zφ,Jφ=J2φ,JZφ=φ,ZJφ.\displaystyle\langle\varphi,ZJ\varphi\rangle=\langle Z^{*}\varphi,J\varphi\rangle=\langle J^{2}\varphi,JZ^{*}\varphi\rangle=-\langle\varphi,ZJ\varphi\rangle\,.

Thus φ,φ~=0\langle\varphi,\widetilde{\varphi}\rangle=0. Also

(A.1) ZZφ~=ZZZJφ=ZJZZφ=ZJλφ=λφ~.\displaystyle ZZ^{*}\widetilde{\varphi}=ZZ^{*}ZJ\varphi=ZJZZ^{*}\varphi=ZJ\lambda\varphi=\lambda\widetilde{\varphi}\,.

Thus φ~E\widetilde{\varphi}\in E, and moreover, we have

ZJφ~=ZJZJ=J2ZZφ=λφ.\displaystyle ZJ\widetilde{\varphi}=ZJZJ=J^{2}ZZ^{*}\varphi=-\lambda\varphi\,.

Thus the span of {φ,φ~}\Set{\varphi,\widetilde{\varphi}} is invariant under the action of ZJZJ.

Pick ψ\psi in the orthogonal of the span of {φ,φ~}\Set{\varphi,\widetilde{\varphi}} in EE. We can form ψ~:=ZJψ\widetilde{\psi}:=ZJ\psi similar as before, where we have ψ,ψ~=0\langle\psi,\widetilde{\psi}\rangle=0 and ψ~E\widetilde{\psi}\in E. In particular η,ψ~=0\langle\eta,\widetilde{\psi}\rangle=0 for η\eta in the span of {φ,φ~}\Set{\varphi,\widetilde{\varphi}} since

η,ψ~=η,ZJψ=Zη,Jψ=J2ψ,JZη=ψ,ZJη=0.\displaystyle\langle\eta,\widetilde{\psi}\rangle=\langle\eta,ZJ\psi\rangle=\langle Z^{*}\eta,J\psi\rangle=\langle J^{2}\psi,JZ^{*}\eta\rangle=-\langle\psi,ZJ\eta\rangle=0\,.

Thus the eigen-subspace EE is even-dimensional. This implies that imZ\operatorname{im}Z is even-dimensional.

We may also recast the above theorem somewhat differently as follows

Theorem A.3.

If FZi()F\in Z\mathcal{F}_{\operatorname{i}\mathbb{R}}\left(\mathcal{H}\right) as in Section˜2, i.e., FF is a self-adjoint Fredholm with CC is a real structure on \mathcal{H}, such that

{F,C}=FC+CF=0\displaystyle\Set{F,C}=FC+CF=0

then

ind2(F)dim(ker(F))mod22\displaystyle\operatorname{ind}_{2}\left(F\right)\equiv\dim\left(\ker\left(F\right)\right)\mod 2\in\mathbb{Z}_{2}

is well-defined, in the sense that if GZi()G\in Z\mathcal{F}_{\operatorname{i}\mathbb{R}}\left(\mathcal{H}\right) and FG\left\lVert F-G\right\rVert is sufficiently small, then

ind2(G)=ind2(F).\displaystyle\operatorname{ind}_{2}(G)=\operatorname{ind}_{2}(F)\,.
Proof.

Since FF is self-adjoint, then kerF=(imF)=:A\ker F=(\operatorname{im}F)^{\perp}=:A and (kerF)=imF=:B(\ker F)^{\perp}=\operatorname{im}F=:B. Decompose GG in ABA\oplus B, we write

G=[GAAGABGABGBB]\displaystyle G=\begin{bmatrix}G_{AA}&G_{AB}\\ G_{AB}^{*}&G_{BB}\end{bmatrix}

Since F:(kerF)imFF:(\ker F)^{\perp}\to\operatorname{im}F is invertible and GF\left\lVert G-F\right\rVert small, then GBBG_{BB} is invertible. Define the Schur complement Z:AAZ:A\to A as

Z=GAAGABGBB1GAB\displaystyle Z=G_{AA}-G_{AB}G_{BB}^{-1}G_{AB}^{*}

Since GG is self-adjoint, then ZZ is, too. Since GC=CGGC=-CG, then the subspaces A,BA,B are both invariant under the action of CC. Thus the expressions

GAAC=CGAA,GBBC=CBB,GABC=CGAB\displaystyle G_{AA}C=-CG_{AA},\quad G_{BB}C=-C_{BB},\quad G_{AB}C=-CG_{AB}

make sense and hold. It follows that

ZC=CZ.\displaystyle ZC=-CZ\,.

Similar to Lemma˜A.1, one also has

dim(kerG)=dim(kerF)dim(imZ).\displaystyle\dim(\ker G)=\dim(\ker F)-\dim(\operatorname{im}Z)\,.

We argue that imZ\operatorname{im}Z is even-dimensional. Since ZZ is self-adjoint, (kerZ)=imZ=:V(\ker Z)^{\perp}=\operatorname{im}Z=:V which is finite-dimensional. View Z:VVZ:V\to V as invertible operator. Since Z2Z^{2} is self-adjoint, the space VV admits an eigen-subspace decomposition with respect to Z2Z^{2}. Let EE be one of the eigen-subspace and pick φE\varphi\in E where Z2φ=λφZ^{2}\varphi=\lambda\varphi. Let φ~:=ZCφ\widetilde{\varphi}:=ZC\varphi. Note CφVC\varphi\in V since ZZ is CC-real. Thus φ~\widetilde{\varphi} is well-defined. Now

φ,φ~=φ,ZCφ=Zφ,Cφ=φ,CZφ=φ,ZCφ=φ,φ~\displaystyle\langle\varphi,\widetilde{\varphi}\rangle=\langle\varphi,ZC\varphi\rangle=\langle Z\varphi,C\varphi\rangle=\langle\varphi,CZ\varphi\rangle=-\langle\varphi,ZC\varphi\rangle=-\langle\varphi,\widetilde{\varphi}\rangle

and hence φ,φ~=0\langle\varphi,\widetilde{\varphi}\rangle=0. Also

Z2φ~=Z2(ZCφ)=ZCZ2φ=λZCφ=λφ~.\displaystyle Z^{2}\widetilde{\varphi}=Z^{2}(ZC\varphi)=ZCZ^{2}\varphi=\lambda ZC\varphi=\lambda\widetilde{\varphi}\,.

Thus φ~E\widetilde{\varphi}\in E. Moreover, we have

ZCφ~=ZCZCφ=C2Z2φ=λφ\displaystyle ZC\widetilde{\varphi}=ZCZC\varphi=C^{2}Z^{2}\varphi=\lambda\varphi

and hence the span of {φ,φ~}\Set{\varphi,\widetilde{\varphi}} is invariant under the action of ZCZC.

Pick ψ\psi in the orthogonal complement of the span of {φ,φ~}\Set{\varphi,\widetilde{\varphi}} in EE. Similarly construct ψ~=ZCψ\widetilde{\psi}=ZC\psi such that ψ,ψ~=0\langle\psi,\widetilde{\psi}\rangle=0 and ψ~E\widetilde{\psi}\in E. In particular η,ψ~=0\langle\eta,\widetilde{\psi}\rangle=0 for η\eta in the span of {φ,φ~}\Set{\varphi,\widetilde{\varphi}} since

η,ψ~=η,ZCψ=Zη,Cψ=C2ψ,CZη=ψ,ZCη=0.\displaystyle\langle\eta,\widetilde{\psi}\rangle=\langle\eta,ZC\psi\rangle=\langle Z\eta,C\psi\rangle=\langle C^{2}\psi,CZ\eta\rangle=\langle\psi,-ZC\eta\rangle=0\,.

Thus the eigen-subspace EE is even-dimensional. This implies that imZ\operatorname{im}Z is even-dimensional. ∎

Corollary A.4.

The Atiyah-Singer 2\mathbb{Z}_{2} index is stable under symmetric compact perturbations:

  1. 1.

    If FF\in\mathcal{F}_{\star\mathbb{H}} and K𝒦K\in\mathcal{K}_{\star\mathbb{H}}, then ind2(F+K)=ind2F\operatorname{ind}_{2}(F+K)=\operatorname{ind}_{2}F.

  2. 2.

    If FisaF\in\mathcal{F}_{\operatorname{i}\mathbb{R}}^{\mathrm{sa}} and K𝒦isaK\in\mathcal{K}_{\operatorname{i}\mathbb{R}}^{\mathrm{sa}}, then ind2(F+K)=ind2F\operatorname{ind}_{2}(F+K)=\operatorname{ind}_{2}F.

Proof.

Consider a straight-line homotopy from FF to F+KF+K and use Theorem˜A.2 and Theorem˜A.3. ∎

Appendix B A child’s garden of homotopies

In this section we employ the same notational conventions as in Section˜2. We are concerned with homotopies of unitaries and self-adjoint projections without locality constraints.

B.1 Equivariant homotopies of unitaries

The following theorem was presented in [Kui65]. The proof which was outlined in eq.˜1.1 applies only to the case 𝔽=\mathbb{F}=\mathbb{C}, the other two cases may be found in Kuiper’s original paper.

Theorem B.1 (Kuiper).

For any 𝔽{,,}\mathbb{F}\in\Set{\mathbb{C},\mathbb{R},\mathbb{H}} and any invertible operator A𝒢𝔽A\in\mathcal{G}_{\mathbb{F}}, there is a continuous path from AA to 𝟙\mathds{1} which passes within 𝒢𝔽\mathcal{G}_{\mathbb{F}}. If AA is unitary the path passes through unitaries.

New (to us) is the following \star-variant of it:

Theorem B.2.

Let 𝔽{,}\mathbb{F}\in\Set{\star\mathbb{R},\star\mathbb{H}}. Then

π0(𝒰𝔽){0}\displaystyle\pi_{0}(\mathcal{U}_{\mathbb{F}})\simeq\Set{0}
Proof.

Let F{C,J}F\in\Set{C,J}. For bi-variate polynomials p(z,z¯)=αznz¯mp(z,\bar{z})=\alpha z^{n}\bar{z}^{m}, we have

p(U,U)F=αUn(U)mF=α¯(U)nUm=F(p(U,U)).\displaystyle p(U,U^{\ast})F=\alpha U^{n}(U^{\ast})^{m}F=\bar{\alpha}(U^{\ast})^{n}U^{m}=F(p(U,U^{\ast}))^{\ast}\,.

Thus for a continuous function fC(σ(U))f\in C(\sigma(U)), one has f(U)𝔽f(U)\in\mathcal{B}_{\mathbb{F}}.

Consider the square root function

h(z)=exp(iarg(z)/2)(z{0})\displaystyle h(z)=\exp(\operatorname{i}\arg(z)/2)\qquad(z\in\mathbb{C}\setminus\Set{0})

where we take arg(z)[0,2π)\arg(z)\in[0,2\pi) for concreteness. The function hh is bounded measurable on σ(U)\sigma(U), and clearly there exists a sequence of continuous functions fnf_{n} on σ(U)\sigma(U) that converges point-wise to hh, and fn\|f_{n}\|_{\infty} is bounded. By the spectral theorem [RS80, Theorem VII.2(d)], fn(U)f_{n}(U) converges strongly to h(U)h(U). Thus h(U)𝔽h(U)\in\mathcal{B}_{\mathbb{F}}. In particular, since h¯(z)h(z)=1\bar{h}(z)h(z)=1, then h(U)𝒰𝔽h(U)\in\mathcal{U}_{\mathbb{F}}. Write h(U)=Fh(U)Fh(U)=Fh(U)^{*}F^{*}, then

U=h(U)2=h(U)Fh(U)F.\displaystyle U=h(U)^{2}=h(U)Fh(U)^{*}F^{*}\,.

Use Theorem˜B.1 to construct a continuous path of unitaries [0,1]tVt[0,1]\ni t\mapsto V_{t} connecting h(U)h(U) to 𝟙\mathds{1}, we let Ut=VtFVtFU_{t}=V_{t}FV_{t}^{*}F^{*}. Then

UtF=VtFVt=F(FVtFVt)=FUt\displaystyle U_{t}F=V_{t}FV_{t}^{*}=F(FV_{t}F^{*}V_{t}^{*})=FU_{t}^{*}

Thus Ut𝒰𝔽U_{t}\in\mathcal{U}_{\mathbb{F}} and connects UU and 𝟙\mathds{1}. ∎

We will make use of the fact that the polar part preserves symmetry constraints:

Lemma B.3.

Let 𝔽{,,,}\mathbb{F}\in\Set{\mathbb{R},\mathbb{H},\star\mathbb{R},\star\mathbb{H}}. If A𝔽A\in\mathcal{B}_{\mathbb{F}}, then pol(A)𝔽\operatorname{pol}(A)\in\mathcal{B}_{\mathbb{F}}.

Proof.

Let F{C,J}F\in\Set{C,J} according to 𝔽\mathbb{F}.

First consider the cases 𝔽{,}\mathbb{F}\in\Set{\mathbb{R},\mathbb{H}}. Since 𝔽\mathcal{B}_{\mathbb{F}} is closed under the adjoint operation, |A|2𝔽|A|^{2}\in\mathcal{B}_{\mathbb{F}} too, and hence the polar part, by writing it as the strong limit of functions which approximate A|A|1A|A|^{-1}. Indeed, pol(A)=slimAfn(|A|)\operatorname{pol}(A)=\operatorname*{s-lim}Af_{n}(|A|) where fn(λ)=1/λf_{n}(\lambda)=1/\lambda if λ1/n\lambda\geq 1/n and fn(λ)=1/nf_{n}(\lambda)=1/n if λ1/n\lambda\leq 1/n. In particular, fn:f_{n}:\mathbb{R}\to\mathbb{R} is \mathbb{R}-valued and hence fn(|A|)F=Ffn(|A|)f_{n}(|A|)F=Ff_{n}(|A|). Then pol(A)F=slimAfn(|A|)F=FslimAfn(|A|)=Fpol(A)\operatorname{pol}(A)F=\operatorname*{s-lim}Af_{n}(|A|)F=F\operatorname*{s-lim}Af_{n}(|A|)=F\operatorname{pol}(A).

Next, for the case 𝔽{,}\mathbb{F}\in\Set{\star\mathbb{R},\star\mathbb{H}}, we have

AAF=AFA=FAA.\displaystyle A^{*}AF=A^{*}FA^{*}=FAA^{*}.

Then fn(|A|)F=Ffn(|A|)f_{n}(|A|)F=Ff_{n}(|A^{*}|) and

pol(A)F=slimAfn(|A|)F=FslimAfn(|A|)=Fpol(A)=F(pol(A)).\displaystyle\operatorname{pol}(A)F=\operatorname*{s-lim}Af_{n}(|A|)F=F\operatorname*{s-lim}A^{*}f_{n}(|A^{*}|)=F\operatorname{pol}(A^{*})=F(\operatorname{pol}(A))^{*}.

Lemma B.4.

For 𝔽=,,\mathbb{F}=\mathbb{C},\mathbb{R},\mathbb{H}, let Z𝔽Z\in\mathcal{F}_{\mathbb{F}} be essentially unitary with zero Fredholm index. Then there is a unitary operator Y𝒰𝔽Y\in\mathcal{U}_{\mathbb{F}} such that YZ𝒦Y-Z\in\mathcal{K}.

Proof.

In what follows, we let F=𝟙,C,JF=\mathds{1},C,J according to the value of 𝔽\mathbb{F}. Using the fact that T=TT=T^{\ast}\in\mathcal{B} has T𝒦T\in\mathcal{K} iff σess(T)={0}\sigma_{\mathrm{ess}}(T)=\Set{0}, since 𝟙|Z|2\mathds{1}-|Z|^{2} is compact, its essential spectrum equals {1}\Set{1} which implies 𝟙|Z|\mathds{1}-|Z| is compact as well. Let pol(Z)\operatorname{pol}(Z) denote the polar part of ZZ. Then by the above,

(B.1) Zpol(Z)=pol(Z)|Z|pol(Z)=pol(Z)(|Z|𝟙)𝒦.\displaystyle Z-\operatorname{pol}(Z)=\operatorname{pol}(Z)|Z|-\operatorname{pol}(Z)=\operatorname{pol}(Z)(|Z|-\mathds{1})\in\mathcal{K}\,.

Now, since indZ=0\operatorname{ind}Z=0, kerZ\ker Z and (imZ)(\operatorname{im}Z)^{\perp} are finite-dimensional and of the same dimension, we let M:kerZim(Z)M:\ker Z\to\operatorname{im}(Z)^{\perp} be any unitary map between two finite vector spaces of the same dimension and define Y:=pol(Z)MY:=\operatorname{pol}(Z)\oplus M which is now unitary and YZY-Z is compact using eq.˜B.1 and the fact MM is finite rank. This settles the case 𝔽=\mathbb{F}=\mathbb{C}.

Next, if F𝟙F\neq\mathds{1}, we have pol(Z)F=Fpol(Z)\operatorname{pol}(Z)F=F\operatorname{pol}(Z) too using Lemma˜B.3 and ZF=FZZF=FZ implies

FkerZ=kerZ,F(imZ)=(imZ).\displaystyle F\ker Z=\ker Z,\quad F(\operatorname{im}Z)^{\perp}=(\operatorname{im}Z)^{\perp}\,.

Now the analysis divides according to the value of 𝔽\mathbb{F}. When 𝔽=\mathbb{F}=\mathbb{R}, we have from Lemma˜B.5 right below bases {φi}i=1m\Set{\varphi_{i}}_{i=1}^{m} and {ψi}i=1m\Set{\psi_{i}}_{i=1}^{m} for kerZ\ker Z and (imZ)(\operatorname{im}Z)^{\perp}, respectively, such that φi,ψi\varphi_{i},\psi_{i} are fixed by CC. Let M:kerZ(imZ)M:\ker Z\to(\operatorname{im}Z)^{\perp} be the unitary operator mapping φiψi\varphi_{i}\mapsto\psi_{i} for i=1,,mi=1,\dots,m. Then

CMφi=Cψi=ψi=Mφi=MCφi.\displaystyle CM\varphi_{i}=C\psi_{i}=\psi_{i}=M\varphi_{i}=MC\varphi_{i}\,.

Thus the unitary direct sum Y:=pol(Z)MY:=\operatorname{pol}(Z)\oplus M commutes with CC.

When 𝔽=\mathbb{F}=\mathbb{H}, applying Lemma˜B.5 again, we obtain bases of Kramers pairs {φi,φi+m}i=1m\Set{\varphi_{i},\varphi_{i+m}}_{i=1}^{m} and {ψi,ψi+m}i=1m\Set{\psi_{i},\psi_{i+m}}_{i=1}^{m} for kerZ\ker Z and (imZ)(\operatorname{im}Z)^{\perp}, respectively where mm is half the dimension of the kernel. Let M:kerZ(imZ)M:\ker Z\to(\operatorname{im}Z)^{\perp} be defined as

φiψi+m,φi+mψi.\displaystyle\varphi_{i}\mapsto-\psi_{i+m},\quad\varphi_{i+m}\mapsto\psi_{i}\,.

Then

JMφi=Jψi+m=ψi=Mφi+m=MJφi\displaystyle JM\varphi_{i}=-J\psi_{i+m}=\psi_{i}=M\varphi_{i+m}=MJ\varphi_{i}

and similarly for JMφi+m=MJφi+mJM\varphi_{i+m}=MJ\varphi_{i+m}.

We conclude that in all three cases, one extends pol(Z)\operatorname{pol}(Z) to a unitary operator Y:=pol(Z)MY:=\operatorname{pol}(Z)\oplus M with [F,Y]=0[F,Y]=0 so Y𝒰𝔽Y\in\mathcal{U}_{\mathbb{F}} and moreover, YZ𝒦Y-Z\in\mathcal{K}. ∎

Above we have used the following equivariant basis assertion:

Lemma B.5.

Let VV be a Hilbert space with dimension mm, possibly infinite.

  1. 1.

    Suppose there is an anti-unitary C:VVC:V\to V with C2=𝟙C^{2}=\mathds{1}. Then VV has an ONB {φi}i=1m\Set{\varphi_{i}}_{i=1}^{m} such that Cφi=φiC\varphi_{i}=\varphi_{i}. If m=2lm=2l for some l{}l\in\mathbb{N}\cup\Set{\infty}, then VV has an ONB {φi,ψi}i=1l\Set{\varphi_{i},\psi_{i}}_{i=1}^{l} such that Cφi=ψiC\varphi_{i}=\psi_{i}.

  2. 2.

    If there is an anti-unitary J:VVJ:V\to V with J2=𝟙J^{2}=-\mathds{1}, VV has an orthonormal basis consisting of Kramers pairs, i.e., there is an ONB {φi,ψi}i=1m\Set{\varphi_{i},\psi_{i}}_{i=1}^{m} such that Jφi=ψiJ\varphi_{i}=\psi_{i}. In particular, if dimV<\dim V<\infty, it follows that dimV=2m.\dim V=2m.

Proof.

The first part is [GP06, Lemma 1] which we reproduce here for completeness. Consider the subset W={φ+Cφ|φV}W=\Set{\varphi+C\varphi}{\varphi\in V} consisting of elements from VV. The elements in WW are fixed by CC, and WW is an \mathbb{R}-vector space. To verify this, let α\alpha\in\mathbb{R} and ψ=φ+CφW\psi=\varphi+C\varphi\in W, then αψ=αφ+C(αφ)W\alpha\psi=\alpha\varphi+C(\alpha\varphi)\in W. Let {φi}i=1q\Set{\varphi_{i}}_{i=1}^{q} be an orthonormal basis for WW considered as an \mathbb{R}-vector space. Here qq can be finite or infinite. Now, for any ψV\psi\in V, let η:=iψ\eta:=-\operatorname{i}\psi, and we have the identity

ψ=12((ψ+Cψ)+i(η+Cη)).\displaystyle\psi=\frac{1}{2}((\psi+C\psi)+\operatorname{i}(\eta+C\eta))\,.

By expanding now ψ+Cψ\psi+C\psi and η+Cη\eta+C\eta in the \mathbb{R}-basis of WW, we conclude that any element ψV\psi\in V may be written as a \mathbb{C}-linear combination of {φi}i=1q\Set{\varphi_{i}}_{i=1}^{q}. In fact the set {φi}i=1q\Set{\varphi_{i}}_{i=1}^{q} is orthonormal within VV, since WW inherits the same inner product structure with which {φi}i=1q\Set{\varphi_{i}}_{i=1}^{q} is orthonormal. Thus {φi}i=1q\Set{\varphi_{i}}_{i=1}^{q} is an ONB for VV and q=mq=m.

If m=2lm=2l for some l{}l\in\mathbb{N}\cup\Set{\infty}, define

ηj±:=φ2j1±iφ2jφ2j1±iφ2j(j1,,l).\displaystyle\eta_{j}^{\pm}:=\frac{\varphi_{2j-1}\pm\operatorname{i}\varphi_{2j}}{\left\lVert\varphi_{2j-1}\pm\operatorname{i}\varphi_{2j}\right\rVert}\qquad(j\in{1,\dots,l})\,.

Thus {ηj±}j=1l\Set{\eta^{\pm}_{j}}_{j=1}^{l} is an ONB for VV with Cηj±=ηjC\eta_{j}^{\pm}=\eta_{j}^{\mp}.

For the second part, let φ1\varphi_{1} denote a unit-length vector from VV. Let ψ1:=Jφ1\psi_{1}:=J\varphi_{1}. By anti-unitarity, ψ1=1\left\lVert\psi_{1}\right\rVert=1 and by J2=𝟙J^{2}=-\mathds{1},

φ1,ψ1=φ1,Jφ1=J2φ1,Jφ1=φ1,ψ1\displaystyle\langle\varphi_{1},\psi_{1}\rangle=\langle\varphi_{1},J\varphi_{1}\rangle=\langle J^{2}\varphi_{1},J\varphi_{1}\rangle=-\langle\varphi_{1},\psi_{1}\rangle

which implies φ1,ψ1=0\langle\varphi_{1},\psi_{1}\rangle=0. In particular, one has Jψ1=φ1J\psi_{1}=-\varphi_{1} and the span of {φ1,ψ1}\Set{\varphi_{1},\psi_{1}} is invariant under JJ. Pick another φ2\varphi_{2} in the orthogonal complement of the span of {φ1,ψ1}\Set{\varphi_{1},\psi_{1}} and let ψ2=Jφ2\psi_{2}=J\varphi_{2}. One readily verifies that ψ2\psi_{2} is also orthogonal to the span of {φ1,ψ1}\Set{\varphi_{1},\psi_{1}}. Continue until VV is spanned. This construction works for the case when m<m<\infty. When m=m=\infty, we perform a so-called Zornication. ∎

Lemma B.6.

For 𝔽=,\mathbb{F}=\star\mathbb{R},\star\mathbb{H}, let Z𝔽Z\in\mathcal{F}_{\mathbb{F}} be essentially unitary with zero Fredholm index, if applicable. Then there is a unitary Y𝒰𝔽Y\in\mathcal{U}_{\mathbb{F}} such that ZYZ-Y is compact.

Proof.

We have again pol(Z)𝔽\operatorname{pol}(Z)\in\mathcal{B}_{\mathbb{F}} by Lemma˜B.3. To extend pol(Z)\operatorname{pol}(Z) to a unitary operator in 𝒰𝔽\mathcal{U}_{\mathbb{F}}, the analysis divides according to the value of 𝔽\mathbb{F}.

For 𝔽=\mathbb{F}=\star\mathbb{R}, let kerZ\ker Z be spanned by an orthonormal basis {φi}i=1m\Set{\varphi_{i}}_{i=1}^{m}. We note that Z𝔽Z\in\mathcal{B}_{\mathbb{F}} implies

CkerZ=ker(Z)=(imZ).\displaystyle C\ker Z=\ker(Z^{\ast})=(\operatorname{im}Z)^{\perp}\,.

Thus (imZ)(\operatorname{im}Z)^{\perp} is spanned by the orthonormal basis {Cφi}i=1m\Set{C\varphi_{i}}_{i=1}^{m}. Let M:kerZ(imZ)M:\ker Z\to(\operatorname{im}Z)^{\perp} be the unitary which maps φi\varphi_{i} to CφiC\varphi_{i}. Then

CMφi=CCφi=φi=MCφ\displaystyle CM\varphi_{i}=CC\varphi_{i}=\varphi_{i}=M^{*}C\varphi

and we may extend pol(Z)\operatorname{pol}(Z) by MM then.

Next, consider the case 𝔽=\mathbb{F}=\star\mathbb{H}. Since the index is zero, dimkerZ2\dim\ker Z\in 2\mathbb{N}, so let {φi,φi+m}i=1m\Set{\varphi_{i},\varphi_{i+m}}_{i=1}^{m} denote an orthonormal basis. In this case, the fact JkerZ=(imZ)J\ker Z=(\operatorname{im}Z)^{\perp} still holds. Thus, (imZ)(\operatorname{im}Z)^{\perp} is spanned by the orthonormal basis {Jφi,Jφi+m}i=1m\Set{J\varphi_{i},J\varphi_{i+m}}_{i=1}^{m}. Define the unitary map M:kerZ(imZ)M:\ker Z\to(\operatorname{im}Z)^{\perp} via

φiJφi+m,φi+mJφi.\displaystyle\varphi_{i}\mapsto-J\varphi_{i+m},\quad\varphi_{i+m}\mapsto J\varphi_{i}\,.

Then

JMφi=J2φi+m=φi+m=MJφi\displaystyle JM\varphi_{i}=-J^{2}\varphi_{i+m}=\varphi_{i+m}=M^{*}J\varphi_{i}

and similarly JMφi+m=MJφi+mJM\varphi_{i+m}=M^{*}J\varphi_{i+m}. So again we define Y:=pol(Z)MY:=\operatorname{pol}(Z)\oplus M and Y𝒰𝔽Y\in\mathcal{U}_{\mathbb{F}}. ∎

B.2 Equivariant homotopies of self-adjoint unitaries

Lemma B.7.

Let 𝔽{,,,i,i}\mathbb{F}\in\Set{\mathbb{C},\mathbb{R},\mathbb{H},\operatorname{i}\mathbb{R},\operatorname{i}\mathbb{H}}. The non-trivial SAUs in 𝒮𝔽\mathcal{S}_{\mathbb{F}} are nullhomotopic.

Proof.

Let U,U~𝒮𝔽U,\widetilde{U}\in\mathcal{S}_{\mathbb{F}} be non-trivial. In the case when 𝔽=\mathbb{F}=\mathbb{C}, we can simply choose a unitary operator WW that maps the ±1\pm 1 eigenspaces of UU to the respective ±1\pm 1 eigenspaces of U~\widetilde{U}, since these eigenspaces are infinite dimensional by the non-triviality assumption. It is clear that

(B.2) U=WU~W.\displaystyle U=W^{*}\widetilde{U}W\,.

We apply Theorem˜B.1 to deform Wt𝟙W_{t}\rightsquigarrow\mathds{1} withing 𝒰\mathcal{U}_{\mathbb{C}} and obtain the desired path WtU~WtUW_{t}^{*}\widetilde{U}W_{t}\rightsquigarrow U within 𝒮.\mathcal{S}_{\mathbb{C}}.

We consider 𝔽=\mathbb{F}=\mathbb{R}. Since UC=CUUC=CU, it follows that

Cker(U±𝟙)=ker(U±𝟙).\displaystyle C\ker(U\pm\mathds{1})=\ker(U\pm\mathds{1})\,.

Thus, we apply Lemma˜B.5 to obtain an ONB {φi±}i=1\Set{\varphi_{i}^{\pm}}_{i=1}^{\infty} for the ±1\pm 1 eigenspace of UU such that Cφi±=φi±C\varphi_{i}^{\pm}=\varphi_{i}^{\pm}. Similarly, let {φ~i±}i=1\Set{\widetilde{\varphi}_{i}^{\pm}}_{i=1}^{\infty} be an ONB fixed by CC for the ±1\pm 1 eigenspaces of U~\widetilde{U}. Let WW be the unitary operator that maps φi±φ~i±\varphi_{i}^{\pm}\mapsto\widetilde{\varphi}_{i}^{\pm}. It follows that eq.˜B.2 holds and, moreover, WC=CWWC=CW holds. Indeed, we have

WCφi±=Wφi±=φ~i±=Cφ~i±=CWφi±.\displaystyle WC\varphi_{i}^{\pm}=W\varphi_{i}^{\pm}=\widetilde{\varphi}_{i}^{\pm}=C\widetilde{\varphi}_{i}^{\pm}=CW\varphi_{i}^{\pm}\,.

Thus, we can deform Wt𝟙W_{t}\rightsquigarrow\mathds{1} within 𝒰\mathcal{U}_{\mathbb{R}} using Theorem˜B.1, and this gives the desired path WtU~WtUW_{t}^{*}\widetilde{U}W_{t}\rightsquigarrow U within 𝒮\mathcal{S}_{\mathbb{R}}. The 𝔽=\mathbb{F}=\mathbb{H} case is similar to 𝔽=\mathbb{F}=\mathbb{R}. Using the fact that Jker(U±𝟙)=ker(U±𝟙)J\ker(U\pm\mathds{1})=\ker(U\pm\mathds{1}) from UJ=JUUJ=JU and Lemma˜B.5, we find an ONB {φi±,ψi±}i=1\Set{\varphi^{\pm}_{i},\psi^{\pm}_{i}}_{i=1}^{\infty} for the ±1\pm 1 eigenspaces of UU with Jφi±=ψi±J\varphi_{i}^{\pm}=\psi_{i}^{\pm}, and an ONB {φ~i±,ψ~i±}i=1\Set{\widetilde{\varphi}^{\pm}_{i},\widetilde{\psi}^{\pm}_{i}}_{i=1}^{\infty} for U~\widetilde{U} similarly constructed. We let WW maps φi±,ψi±\varphi^{\pm}_{i},\psi^{\pm}_{i} to φ~i±,ψ~i±\widetilde{\varphi}^{\pm}_{i},\widetilde{\psi}^{\pm}_{i} respectively. Then eq.˜B.2 holds, and so does WJ=JWWJ=JW. Indeed, we have

WJφi±=Wψi±=ψ~i±=Jφ~i±=JWφi±\displaystyle WJ\varphi^{\pm}_{i}=W\psi^{\pm}_{i}=\widetilde{\psi}^{\pm}_{i}=J\widetilde{\varphi}^{\pm}_{i}=JW\varphi_{i}^{\pm}
WJψi±=Wφi±=φ~i±=Jψ~i±=JWψi±.\displaystyle WJ\psi^{\pm}_{i}=-W\varphi^{\pm}_{i}=-\widetilde{\varphi}^{\pm}_{i}=J\widetilde{\psi}^{\pm}_{i}=JW\psi_{i}^{\pm}\,.

Similar to the previous case, it follows that there exists a path UU~U\rightsquigarrow\widetilde{U} within 𝒮\mathcal{S}_{\mathbb{H}} using the conjugate operator WW.

We turn to consider 𝔽=i,i\mathbb{F}=\operatorname{i}\mathbb{R},\operatorname{i}\mathbb{H}. Let FF denote CC or JJ. Since UF=FUUF=-FU, it follows that

Fker(U±𝟙)=ker(U𝟙).\displaystyle F\ker(U\pm\mathds{1})=\ker(U\mp\mathds{1})\,.

Let {φi+}i=1\Set{\varphi_{i}^{+}}_{i=1}^{\infty} be an ONB for the +1+1 eigenspace of UU. Using the above relation, then {φi:=Fφi+}i=1\Set{\varphi_{i}^{-}:=F\varphi_{i}^{+}}_{i=1}^{\infty} is an ONB for the 1-1 eigenspace of UU. Let {φ~i+}i=1\Set{\widetilde{\varphi}_{i}^{+}}_{i=1}^{\infty} and {φ~i:=Fφ~i+}i=1\Set{\widetilde{\varphi}_{i}^{-}:=F\widetilde{\varphi}_{i}^{+}}_{i=1}^{\infty} be a similar construction of ONB for the ±1\pm 1 eigenspaces of U~\widetilde{U}. Define WW that maps φi±φ~i±\varphi_{i}^{\pm}\mapsto\widetilde{\varphi}_{i}^{\pm}. Then eq.˜B.2 holds and, moreover, WF=FWWF=FW holds. Indeed, we have

WFφi+=Wφi=φ~i=Fφ~i+=FWφi+\displaystyle WF\varphi_{i}^{+}=W\varphi_{i}^{-}=\widetilde{\varphi}_{i}^{-}=F\widetilde{\varphi}_{i}^{+}=FW\varphi_{i}^{+}
WFφi=WFFφi+=±Wφi+=±φ~i+=Fφ~i=FWφi\displaystyle WF\varphi_{i}^{-}=WFF\varphi_{i}^{+}=\pm W\varphi_{i}^{+}=\pm\widetilde{\varphi}_{i}^{+}=F\widetilde{\varphi}_{i}^{-}=FW\varphi_{i}^{-}

where in the last line the +1,1+1,-1 prefactors correspond to F=C,JF=C,J respectively. We then use Theorem˜B.1 to deform Wt𝟙W_{t}\rightsquigarrow\mathds{1} within 𝒰𝔽\mathcal{U}_{\mathbb{F}}, and the path Ut:=WtU~WtUU_{t}:=W_{t}^{*}\widetilde{U}W_{t}\rightsquigarrow U will be within 𝒮𝔽\mathcal{S}_{\mathbb{F}}. Indeed, UtU_{t} is a SAU and UtF=WtU~WtF=FWtU~Wt=FUtU_{t}F=W_{t}^{*}\widetilde{U}W_{t}F=-FW_{t}^{*}\widetilde{U}W_{t}=-FU_{t}.

Lemma B.8.

Let PP be essentially a projection in the sense that P2P𝒦P^{2}-P\in\mathcal{K} and PP𝒦P^{*}-P\in\mathcal{K}. Then there exists a self-adjoint projection QQ such that PQ𝒦P-Q\in\mathcal{K}. If PP commutes with a given anti-unitary FF, then so does QQ.

Proof.

Let P~=12(P+P)=P+12(PP)\widetilde{P}=\frac{1}{2}(P+P^{*})=P+\frac{1}{2}(P^{*}-P), then P~\widetilde{P} is self-adjoint and P~2P~𝒦\widetilde{P}^{2}-\widetilde{P}\in\mathcal{K}. Therefore, WLOG we assume PP is self-adjoint. Since P2PP^{2}-P is self-adjoint and compact, its spectrum can only accumulate at 0. Thus the spectrum of PP can only accumulate at 0 and +1+1. Pick any λ0(0,1)σ(P)\lambda_{0}\in(0,1)\setminus\sigma(P). Consider the self-adjoint projection Q=χ(λ0,)(P)Q=\chi_{(\lambda_{0},\infty)}(P). Now

σ(PQ)={λχ(λ0,)(λ)|λσ(P)}.\displaystyle\sigma(P-Q)=\Set{\lambda-\chi_{(\lambda_{0},\infty)}(\lambda)}{\lambda\in\sigma(P)}\,.

Thus the spectrum of PQP-Q can only accumulates at 0, and hence PQ𝒦P-Q\in\mathcal{K}.

Finally, since we pick λ0\lambda_{0} outside the spectrum χ(λ0,)\chi_{(\lambda_{0},\infty)} is a continuous function which may be approximated uniformly (as PP is bounded) by a sequence of polynomials with real coefficients, and hence, we may guarantee that QQ commutes with the anti-unitary FF as well. ∎

Appendix C Stummel idempotents

To study the space of operators of the form A+BRA+BR in Lemma˜5.18, we will construct a suitable grading for these operators, using so-called Stummel idempotents. To motivate the construction, we recall the more familiar concept of the Riesz projections–actually they are only self-adjoint if the associated operator is, otherwise they are merely idempotents which is how we shall refer to them henceforth. They concern the decomposition of operators corresponding to disjoint parts of the spectrum. Let AA be a bounded operator whose spectrum is the disjoint union of two closed subsets E,FE,F of σ(A)\sigma(A). Let us specify even more, so we consider E𝔻E\subset\mathbb{D} and F𝔻¯F\subset\mathbb{C}\setminus\overline{\mathbb{D}} and λ𝟙A\lambda\mathds{1}-A is invertible for λ𝕊1\lambda\in\mathbb{S}^{1}, i.e., there are two parts in the spectrum of AA that are separated by the unit circle. We recall the statement of the theorem concerning Riesz idempotents (see e.g. [GK88, Chapter I]):

Theorem C.1 (Riesz).

The operator

P=12πi𝕊1(λ𝟙A)1dλ\displaystyle P=\frac{1}{2\pi\operatorname{i}}\oint_{\mathbb{S}^{1}}(\lambda\mathds{1}-A)^{-1}\operatorname{d}{\lambda}

is an idempotent such that AA decomposes as

A=[A1100A22]:kerPimPkerPimP.\displaystyle A=\begin{bmatrix}A_{11}&0\\ 0&A_{22}\end{bmatrix}:\ker P\oplus\operatorname{im}P\to\ker P\oplus\operatorname{im}P\,.

In particular, the following operators are invertible

λ𝟙A11\displaystyle\lambda\mathds{1}-A_{11} :kerPkerP(λ𝔻¯)\displaystyle:\ker P\to\ker P\qquad(\lambda\in\overline{\mathbb{D}})
λ𝟙A22\displaystyle\lambda\mathds{1}-A_{22} :imPimP(λ𝔻).\displaystyle:\operatorname{im}P\to\operatorname{im}P\qquad(\lambda\in\mathbb{C}\setminus\mathbb{D})\,.

For the Riesz idempotents, we are concerned with operators of the form A+λ𝟙A+\lambda\mathds{1} (note we switched from considering λ𝟙A\lambda\mathds{1}-A to A+λ𝟙A+\lambda\mathds{1} for notational purposes of the later discussion). The idea can be generalized to operators of the form A+λGA+\lambda G, where A,GA,G are two bounded operators. Instead of considering the spectrum of AA, we will talk about the invertibility of A+λGA+\lambda G for λ\lambda in some subset of the complex plane. For operators of the form A+λGA+\lambda G, we have the following

Theorem C.2.

(Stummel) Let A,GA,G\in\mathcal{B} be given such that Sλ:=A+λGS_{\lambda}:=A+\lambda G is invertible for all λ𝕊1\lambda\in\mathbb{S}^{1}. Then the operators

(C.1) P=12πi𝕊1Sλ1Gdλ,Q=12πi𝕊1GSλ1dλ\displaystyle P=\frac{1}{2\pi\operatorname{i}}\oint_{\mathbb{S}^{1}}S_{\lambda}^{-1}G\operatorname{d}{\lambda},\quad Q=\frac{1}{2\pi\operatorname{i}}\oint_{\mathbb{S}^{1}}GS_{\lambda}^{-1}\operatorname{d}{\lambda}

are idempotents such that with respect to the grading

(C.2) kerPimPkerQimQ\displaystyle\ker P\oplus\operatorname{im}P\to\ker Q\oplus\operatorname{im}Q

the operators A,GA,G decompose diagonally as

(C.3) A=[A1100A22],G=[G1100G22].\displaystyle A=\begin{bmatrix}A_{11}&0\\ 0&A_{22}\end{bmatrix},\quad G=\begin{bmatrix}G_{11}&0\\ 0&G_{22}\end{bmatrix}\,.

Moreover, the following operators are invertible

(C.4) A11+λG11:\displaystyle A_{11}+\lambda G_{11}: kerPkerQ(λ𝔻¯)\displaystyle\ker P\to\ker Q\qquad(\lambda\in\overline{\mathbb{D}})
(C.5) A22+λG22:\displaystyle A_{22}+\lambda G_{22}: imPimQ(λ({})𝔻)\displaystyle\operatorname{im}P\to\operatorname{im}Q\qquad(\lambda\in\left(\mathbb{C}\cup\Set{\infty}\right)\setminus\mathbb{D})

where by λ=\lambda=\infty, we mean the operator G22:imPimQG_{22}:\operatorname{im}P\to\operatorname{im}Q.

We refer the reader to [Stu71] or [GGK13, Chapter IV] for more details and context on the Stummel idempotents. Below we merely reproduce the proof of the convenience of the reader.

Proof of Theorem˜C.2.

The proof is based on the generalized resolvent identity

(C.6) Sλ1Sμ1=(μλ)Sλ1GSμ1.\displaystyle S_{\lambda}^{-1}-S_{\mu}^{-1}=(\mu-\lambda)S_{\lambda}^{-1}GS_{\mu}^{-1}\,.

Define an auxiliary operator

(C.7) K=12πiλ𝕊1Sλ1dλ.\displaystyle K=\frac{1}{2\pi\operatorname{i}}\oint_{\lambda\in\mathbb{S}^{1}}S_{\lambda}^{-1}\operatorname{d}{\lambda}\,.

We first show that

(C.8) KGK=K.\displaystyle KGK=K\,.

There are contours Γ1\Gamma_{1} and Γ2\Gamma_{2} such that Γ2\Gamma_{2} surrounds Γ1\Gamma_{1}, Γ1\Gamma_{1} surrounds 𝕊1\mathbb{S}^{1}, and SλS_{\lambda} is invertible for λ\lambda on these contours. Using the generalized resolvent identity eq.˜C.6, we have

KGK\displaystyle KGK =1(2πi)2Γ1Γ2Sλ1GSμ1dμdλ\displaystyle=\frac{1}{(2\pi\operatorname{i})^{2}}\oint_{\Gamma_{1}}\oint_{\Gamma_{2}}S_{\lambda}^{-1}GS_{\mu}^{-1}\operatorname{d}{\mu}\operatorname{d}{\lambda}
=1(2πi)2Γ1Γ2(μλ)1(Sλ1Sμ1)dμdλ\displaystyle=\frac{1}{(2\pi\operatorname{i})^{2}}\oint_{\Gamma_{1}}\oint_{\Gamma_{2}}(\mu-\lambda)^{-1}(S_{\lambda}^{-1}-S_{\mu}^{-1})\operatorname{d}{\mu}\operatorname{d}{\lambda}
=12πiΓ1[12πiΓ2(μλ)1dμ]Sλ1dλ12πiΓ2[12πiΓ1(μλ)1dλ]Sμ1dμ.\displaystyle=\frac{1}{2\pi\operatorname{i}}\oint_{\Gamma_{1}}\left[\frac{1}{2\pi\operatorname{i}}\oint_{\Gamma_{2}}(\mu-\lambda)^{-1}\operatorname{d}{\mu}\right]S_{\lambda}^{-1}\operatorname{d}{\lambda}-\frac{1}{2\pi\operatorname{i}}\oint_{\Gamma_{2}}\left[\frac{1}{2\pi\operatorname{i}}\oint_{\Gamma_{1}}(\mu-\lambda)^{-1}\operatorname{d}{\lambda}\right]S_{\mu}^{-1}\operatorname{d}{\mu}\,.

Now, for the contour integrals inside the square brackets, the first one is equal to 11 and the second one vanishes, due to the ways we construct the contours, and hence the result. Using eq.˜C.8, we have

P2\displaystyle P^{2} =(KG)2=(KGK)G=KG=P\displaystyle=(KG)^{2}=(KGK)G=KG=P
Q2\displaystyle Q^{2} =(GK)2=G(KGK)=GK=Q.\displaystyle=(GK)^{2}=G(KGK)=GK=Q\,.

Thus PP and QQ are idempotents.

Note the partitions eq.˜C.3 are equivalent to the expressions

AP=QA,GP=QG.\displaystyle AP=QA,\quad GP=QG\,.

Here GP=QGGP=QG readily follows from the construction. For the other one, we need the identity

(C.9) ASλ1G=GSλ1A.\displaystyle AS_{\lambda}^{-1}G=GS_{\lambda}^{-1}A\,.

Indeed, we have

ASλ1G=(A+λG)Sλ1GλGSλ1G=GGSλ1(A+λG)+GSλ1A=GSλ1A.\displaystyle AS_{\lambda}^{-1}G=(A+\lambda G)S_{\lambda}^{-1}G-\lambda GS_{\lambda}^{-1}G=G-GS_{\lambda}^{-1}(A+\lambda G)+GS_{\lambda}^{-1}A=GS_{\lambda}^{-1}A\,.

Using eq.˜C.9, it follows that

AP=12πi𝕊1ASλ1Gdλ=12πi𝕊1GSλ1Adλ=QA.\displaystyle AP=\frac{1}{2\pi\operatorname{i}}\oint_{\mathbb{S}^{1}}AS_{\lambda}^{-1}G\operatorname{d}{\lambda}=\frac{1}{2\pi\operatorname{i}}\oint_{\mathbb{S}^{1}}GS_{\lambda}^{-1}A\operatorname{d}{\lambda}=QA\,.

To show the invertibility for eq.˜C.4 and eq.˜C.5, we construct the inverse operators explicitly. The inverse is Sλ1S_{\lambda}^{-1} for λ𝕊1\lambda\in\mathbb{S}^{1}, and outside the unit circle, naturally we consider

(C.10) Tλ=12πi𝕊1(λμ)1Sμ1dμ(λ𝕊1).\displaystyle T_{\lambda}=\frac{1}{2\pi\operatorname{i}}\oint_{\mathbb{S}^{1}}(\lambda-\mu)^{-1}S_{\mu}^{-1}\operatorname{d}{\mu}\qquad(\lambda\notin\mathbb{S}^{1})\,.

First of all, TλT_{\lambda} decomposes as

Tλ=[(Tλ)1100(Tλ)22]:kerQimQkerPimP\displaystyle T_{\lambda}=\begin{bmatrix}(T_{\lambda})_{11}&0\\ 0&(T_{\lambda})_{22}\end{bmatrix}:\ker Q\oplus\operatorname{im}Q\to\ker P\oplus\operatorname{im}P

so that TλT_{\lambda} can serve as the inverse for SλS_{\lambda} according to the grading eq.˜C.2. To verify the decomposition, it is equivalent to show

TλQ=PTλ(λ𝕊1).\displaystyle T_{\lambda}Q=PT_{\lambda}\qquad(\lambda\notin\mathbb{S}^{1})\,.

Indeed, using the generalized resolvent identity eq.˜C.6, we have

TλQ\displaystyle T_{\lambda}Q =[12πi𝕊1(λμ)1Sμ1dμ][12πi𝕊1GSζ1dζ]\displaystyle=\left[\frac{1}{2\pi\operatorname{i}}\oint_{\mathbb{S}^{1}}(\lambda-\mu)^{-1}S_{\mu}^{-1}\operatorname{d}{\mu}\right]\left[\frac{1}{2\pi\operatorname{i}}\oint_{\mathbb{S}^{1}}GS_{\zeta}^{-1}\operatorname{d}{\zeta}\right]
=1(2πi)2𝕊1𝕊1(λμ)1Sμ1GSζ1dμdζ\displaystyle=\frac{1}{(2\pi\operatorname{i})^{2}}\oint_{\mathbb{S}^{1}}\oint_{\mathbb{S}^{1}}(\lambda-\mu)^{-1}S_{\mu}^{-1}GS_{\zeta}^{-1}\operatorname{d}{\mu}\operatorname{d}{\zeta}
=1(2πi)2𝕊1𝕊1(λμ)1(ζμ)1(Sμ1Sζ1)dμdζ\displaystyle=\frac{1}{(2\pi\operatorname{i})^{2}}\oint_{\mathbb{S}^{1}}\oint_{\mathbb{S}^{1}}(\lambda-\mu)^{-1}(\zeta-\mu)^{-1}(S_{\mu}^{-1}-S_{\zeta}^{-1})\operatorname{d}{\mu}\operatorname{d}{\zeta}
=1(2πi)2𝕊1𝕊1(λμ)1(μζ)1(Sζ1Sμ1)dμdζ\displaystyle=\frac{1}{(2\pi\operatorname{i})^{2}}\oint_{\mathbb{S}^{1}}\oint_{\mathbb{S}^{1}}(\lambda-\mu)^{-1}(\mu-\zeta)^{-1}(S_{\zeta}^{-1}-S_{\mu}^{-1})\operatorname{d}{\mu}\operatorname{d}{\zeta}
=1(2πi)2𝕊1𝕊1(λμ)1Sζ1GSμ1dμdζ\displaystyle=\frac{1}{(2\pi\operatorname{i})^{2}}\oint_{\mathbb{S}^{1}}\oint_{\mathbb{S}^{1}}(\lambda-\mu)^{-1}S_{\zeta}^{-1}GS_{\mu}^{-1}\operatorname{d}{\mu}\operatorname{d}{\zeta}
=[12πi𝕊1Sζ1Gdζ][12πi𝕊1(λμ)1Sμ1dμ]=PTλ.\displaystyle=\left[\frac{1}{2\pi\operatorname{i}}\oint_{\mathbb{S}^{1}}S_{\zeta}^{-1}G\operatorname{d}{\zeta}\right]\left[\frac{1}{2\pi\operatorname{i}}\oint_{\mathbb{S}^{1}}(\lambda-\mu)^{-1}S_{\mu}^{-1}\operatorname{d}{\mu}\right]=PT_{\lambda}\,.

Finally, using

Sλ=(λμ+μ)G+A=(λμ)G+Sμ\displaystyle S_{\lambda}=(\lambda-\mu+\mu)G+A=(\lambda-\mu)G+S_{\mu}

we compute

TλSλ\displaystyle T_{\lambda}S_{\lambda} =12πi𝕊1(λμ)1Sμ1[(λμ)G+Sμ]dμ\displaystyle=\frac{1}{2\pi\operatorname{i}}\oint_{\mathbb{S}^{1}}(\lambda-\mu)^{-1}S_{\mu}^{-1}\left[(\lambda-\mu)G+S_{\mu}\right]\operatorname{d}{\mu}
=12πi𝕊1Sμ1Gdμ12πi𝕊1(μλ)1dμ={Pλ𝔻Pλ𝔻.\displaystyle=\frac{1}{2\pi\operatorname{i}}\oint_{\mathbb{S}^{1}}S_{\mu}^{-1}G\operatorname{d}{\mu}-\frac{1}{2\pi\operatorname{i}}\oint_{\mathbb{S}^{1}}(\mu-\lambda)^{-1}\operatorname{d}{\mu}=\begin{cases}P&\lambda\in\mathbb{D}\\ -P^{\perp}&\lambda\in\mathbb{C}\setminus\mathbb{D}\end{cases}\,.

Similarly, we have

SλTλ={Qλ𝔻Qλ𝔻\displaystyle S_{\lambda}T_{\lambda}=\begin{cases}Q&\lambda\in\mathbb{D}\\ -Q^{\perp}&\lambda\in\mathbb{C}\setminus\mathbb{D}\end{cases}

The invertibility of G22:imPimQG_{22}:\operatorname{im}P\to\operatorname{im}Q heuristically follows by taking λ=\lambda=\infty in eq.˜C.5. In fact, its inverse is exactly the operator eq.˜C.7. This follows from the identities KQ=PKKQ=PK, KG=PKG=P and GK=QGK=Q. Here

KQ=K(GK)=(GK)K=Q.\displaystyle KQ=K(GK)=(GK)K=Q\,.

References

  • [Car26] Élie Cartan “Sur une classe remarquable d’espaces de Riemann” In Bulletin de la Société mathématique de France 54, 1926, pp. 214–264
  • [Car27] Élie Cartan “Sur une classe remarquable d’espaces de Riemann. II” In Bulletin de la Société Mathématique de France 55, 1927, pp. 114–134
  • [Cal41] J.. Calkin “Two-Sided Ideals and Congruences in the Ring of Bounded Operators in Hilbert Space” In Annals of Mathematics 42.4 Annals of Mathematics, 1941, pp. 839–873 URL: http://www.jstor.org/stable/1968771
  • [Dys62] Freeman J Dyson “The threefold way. Algebraic structure of symmetry groups and ensembles in quantum mechanics” In Journal of Mathematical Physics 3.6 American Institute of Physics, 1962, pp. 1199–1215
  • [ABS64] Michael F Atiyah, Raoul Bott and Arnold Shapiro “Clifford modules” In Topology 3 Pergamon, 1964, pp. 3–38
  • [Bar64] Valentine Bargmann “Note on Wigner’s theorem on symmetry operations” In Journal of Mathematical Physics 5.7 American Institute of Physics, 1964, pp. 862–868
  • [Wal64] Charles Terence Clegg Wall “Graded Brauer Groups.” Walter de Gruyter, Berlin/New York Berlin, New York, 1964
  • [Kui65] Nicolaas H. Kuiper “The homotopy type of the unitary group of Hilbert space” In Topology 3.1, 1965, pp. 19–30 DOI: https://doi.org/10.1016/0040-9383(65)90067-4
  • [AS69] Michael F. Atiyah and Isadore M. Singer “Index theory for skew-adjoint Fredholm operators” In Publications Mathématiques de l’IHÉS 37 Institut des Hautes Études Scientifiques, 1969, pp. 5–26 URL: http://www.numdam.org/item/PMIHES_1969__37__5_0/
  • [Stu71] Friedrich Stummel “Diskrete Konvergenz linearer operatoren. II” In Mathematische Zeitschrift 120.3 Springer, 1971, pp. 231–264
  • [Gro73] A. Grothendieck “Topological vector spaces”, Notes on mathematics and its applications New York: GordonBreach, 1973
  • [SSH79] W.. Su, J.. Schrieffer and A.. Heeger “Solitons in Polyacetylene” In Phys. Rev. Lett. 42 American Physical Society, 1979, pp. 1698–1701 DOI: 10.1103/PhysRevLett.42.1698
  • [KDP80] K.. Klitzing, G. Dorda and M. Pepper “New Method for High-Accuracy Determination of the Fine-Structure Constant Based on Quantized Hall Resistance” In Phys. Rev. Lett. 45 American Physical Society, 1980, pp. 494–497 DOI: 10.1103/PhysRevLett.45.494
  • [RS80] Michael Reed and Barry Simon “Methods of Modern Mathematical Physics I: Functional Analysis.” Academic Press Inc., 1980 URL: http://www.amazon.com/Methods-Modern-Mathematical-Physics-Functional/dp/0125850506%3FSubscriptionId%3D0JYN1NVW651KCA56C102%26tag%3Dtechkie-20%26linkCode%3Dxm2%26camp%3D2025%26creative%3D165953%26creativeASIN%3D0125850506
  • [CHO82] AL Carey, CA Hurst and DM O’Brien “Automorphisms of the canonical anticommutation relations and index theory” In Journal of Functional Analysis 48.3 Elsevier, 1982, pp. 360–393
  • [Hal82] PR Halmos “A Hilbert Space Problem Book” Springer Science & Business Media, 1982
  • [Tho+82] D.. Thouless, M. Kohmoto, M.. Nightingale and M. Nijs “Quantized Hall Conductance in a Two-Dimensional Periodic Potential” In Phys. Rev. Lett. 49 American Physical Society, 1982, pp. 405–408 DOI: 10.1103/PhysRevLett.49.405
  • [ASS83] J.. Avron, R. Seiler and B. Simon “Homotopy and Quantization in Condensed Matter Physics” In Phys. Rev. Lett. 51 American Physical Society, 1983, pp. 51–53 DOI: 10.1103/PhysRevLett.51.51
  • [SW86] Barry Simon and Tom Wolff “Singular continuous spectrum under rank one perturbations and localization for random hamiltonians” In Communications on Pure and Applied Mathematics 39.1, 1986, pp. 75–90 DOI: https://doi.org/10.1002/cpa.3160390105
  • [GK88] Israel Gohberg and Marinus A Kaashoek “Block Toeplitz operators with rational symbols” In Contributions to Operator Theory and its Applications Springer, 1988, pp. 385–440
  • [BB89] B. Booss and D.D. Bleecker “Topology and Analysis: The Atiyah-Singer Index Formula and Gauge-Theoretic Physics” Springer, 1989 URL: https://www.amazon.com/Topology-Analysis-Atiyah-Singer-Gauge-Theoretic-Universitext/dp/0387961127?SubscriptionId=0JYN1NVW651KCA56C102&tag=techkie-20&linkCode=xm2&camp=2025&creative=165953&creativeASIN=0387961127
  • [Zak89] J Zak “Berry’s phase for energy bands in solids” In Physical review letters 62.23 APS, 1989, pp. 2747
  • [Mur90] Gerald J. Murphy “C*–Algebras and Operator Theory” Elsevier Inc, 1990 URL: http://gen.lib.rus.ec/book/index.php?md5=2b128a042b434c1d8cf18e0340660c84
  • [BG91] Asher Ben-Artzi and Israel Gohberg “Band matrices and dichotomy” In Topics in matrix and operator Theory Springer, 1991, pp. 137–170
  • [Hat93] Yasuhiro Hatsugai “Chern number and edge states in the integer quantum Hall effect” In Physical review letters 71.22 APS, 1993, pp. 3697
  • [BvS94] J. Bellissard, A. van Elst and H. Schulz-Baldes “The noncommutative geometry of the quantum Hall effect” In J. Math. Phys. 35, 1994, pp. 5373–5451 DOI: 10.1063/1.530758
  • [Con94] A Connes “‘Noncommutative Geometry," Academic Press Inc.” In CA, San Diego, 1994
  • [AZ97] Alexander Altland and Martin R Zirnbauer “Nonstandard symmetry classes in mesoscopic normal-superconducting hybrid structures” In Physical Review B 55.2 APS, 1997, pp. 1142
  • [AG98] M. Aizenman and G.. Graf “Localization bounds for an electron gas” In J. Phys. A Math. Gen. 31, 1998, pp. 6783–6806 DOI: 10.1088/0305-4470/31/32/004
  • [Dou98] Ronald G Douglas “Banach algebra techniques in operator theory”, Graduate texts in mathematics New York, NY: Springer, 1998
  • [Del99] Pierre Deligne “Notes on spinors” In Quantum fields and strings: a course for mathematicians 1.2 AMS, USA, 1999
  • [GVF00] J.M. Gracia-Bondia, J.C. Varilly and H. Figueroa “Elements of Noncommutative Geometry”, Birkhäuser Advanced Texts Basler Lehrbücher Birkhäuser Boston, 2000 URL: https://books.google.com/books?id=2yJIwWbh1isC
  • [Rør+00] Mikael Rørdam, Flemming Larsen, Flemming Larsen and N Laustsen “An introduction to K-theory for C*-algebras” Cambridge University Press, 2000
  • [SKR00] Hermann Schulz-Baldes, Johannes Kellendonk and Thomas Richter “Simultaneous quantization of edge and bulk Hall conductivity” In Journal of Physics A: Mathematical and General 33.2 IOP Publishing, 2000, pp. L27
  • [Kit01] A Yu Kitaev “Unpaired Majorana fermions in quantum wires” In Physics-Uspekhi 44.10S, 2001, pp. 131 DOI: 10.1070/1063-7869/44/10S/S29
  • [EG02] P. Elbau and G.. Graf “Equality of bulk and edge Hall conductance revisited” In Commun. Math. Phys. 229.3, 2002, pp. 415–432 DOI: 10.1007/s00220-002-0698-z
  • [EGS05] A. Elgart, G… Graf and J.H. Schenker “Equality of the bulk and edge Hall conductances in a mobility gap” In Commun. Math. Phys. 259.1, 2005, pp. 185–221 DOI: 10.1007/s00220-005-1369-7
  • [GP06] Stephan Garcia and Mihai Putinar “Complex symmetric operators and applications” In Transactions of the American Mathematical Society 358.3, 2006, pp. 1285–1315
  • [KKR06] Julia Kempe, Alexei Kitaev and Oded Regev “The complexity of the local Hamiltonian problem” In Siam journal on computing 35.5 SIAM, 2006, pp. 1070–1097
  • [Gra07] Gian Michele Graf “Aspects of the Integer Quantum Hall Effect”, 2007
  • [HJS09] Eman Hamza, Alain Joye and Günter Stolz “Dynamical localization for unitary Anderson models” In Mathematical Physics, Analysis and Geometry 12 Springer, 2009, pp. 381–444
  • [Kit09] Alexei Kitaev “Periodic table for topological insulators and superconductors” In AIP Conf. Proc. 1134.1, 2009, pp. 22–30 DOI: http://dx.doi.org/10.1063/1.3149495
  • [HK10] M.. Hasan and C.. Kane “Colloquium: Topological insulators” In Rev. Mod. Phys. 82 American Physical Society, 2010, pp. 3045–3067 DOI: 10.1103/RevModPhys.82.3045
  • [Bae12] John C. Baez “Division Algebras and Quantum Theory” In Foundations of Physics 42.7, 2012, pp. 819–855 DOI: 10.1007/s10701-011-9566-z
  • [FM13] Daniel S. Freed and Gregory W. Moore “Twisted Equivariant Matter” In Annales Henri Poincaré 14.8, 2013, pp. 1927–2023 DOI: 10.1007/s00023-013-0236-x
  • [GGK13] Israel Gohberg, Seymour Goldberg and Marius A Kaashoek “Classes of linear operators” Birkhäuser, 2013
  • [ACL15] Esteban Andruchow, E Chiumiento and ME Di Iorio Lucero “Essentially commuting projections” In Journal of Functional Analysis 268.2 Elsevier, 2015, pp. 336–362
  • [DG15] Giuseppe De Nittis and Kiyonori Gomi “Classification of “Quaternionic" Bloch-Bundles: Topological Quantum Systems of Type AII” In Communications in Mathematical Physics 339 Springer, 2015, pp. 1–55
  • [Sch15] Hermann Schulz-Baldes “Z2-Indices and Factorization Properties of Odd Symmetric Fredholm Operators” In Documenta Mathematica 20, 2015, pp. 1481–1500
  • [BCR16] Chris Bourne, Alan L Carey and Adam Rennie “A non-commutative framework for topological insulators” In Reviews in Mathematical Physics 28.02 World Scientific, 2016, pp. 1650004
  • [GS16] Julian Großmann and Hermann Schulz-Baldes “Index pairings in presence of symmetries with applications to topological insulators” In Communications in Mathematical Physics 343 Springer, 2016, pp. 477–513
  • [KK16] Hosho Katsura and Tohru Koma “The Z2 index of disordered topological insulators with time reversal symmetry” In Journal of Mathematical Physics 57.2 AIP Publishing LLC, 2016, pp. 021903
  • [Kub16] Yosuke Kubota “Notes on twisted equivariant K-theory for C*-algebras” In International Journal of Mathematics 27.06 World Scientific, 2016, pp. 1650058
  • [PS16] Emil Prodan and Hermann Schulz-Baldes “Bulk and Boundary Invariants for Complex Topological Insulators: From K-Theory to Physics” Springer, 2016 DOI: 10.1007/978-3-319-29351-6_4
  • [Thi16] Guo Chuan Thiang “On the K-theoretic classification of topological phases of matter” In Annales Henri Poincaré 17, 2016, pp. 757–794 Springer
  • [THK16] David Thouless, Duncan Haldane and Michael Kosterlitz “Nobel prize for topological pioneers” In Physics Bulletin 29 IOP Publishing Ltd., 2016, pp. 6–7 DOI: 10.1088/2058-7058/29/11/10
  • [BKR17] Chris Bourne, Johannes Kellendonk and Adam Rennie “The K-Theoretic Bulk–Edge Correspondence for Topological Insulators” In Annales Henri Poincaré 18.5, 2017, pp. 1833–1866 DOI: 10.1007/s00023-016-0541-2
  • [GS18] G… Graf and J. Shapiro “The bulk-edge correspondence for disordered chiral chains” In Commun. Math. Phys. 363.3, 2018, pp. 829–846
  • [KK18] Hosho Katsura and Tohru Koma “The noncommutative index theorem and the periodic table for disordered topological insulators and superconductors” 031903 In Journal of Mathematical Physics 59.3, 2018 DOI: 10.1063/1.5026964
  • [Kel19] Johannes Kellendonk “Cyclic Cohomology for Graded C-star algebras and Its Pairings with van Daele K-theory” In Communications in Mathematical Physics 368 Springer, 2019, pp. 467–518
  • [ST19] Jacob Shapiro and Clément Tauber “Strongly Disordered Floquet Topological Systems” In Annales Henri Poincaré 20.6, 2019, pp. 1837–1875 DOI: 10.1007/s00023-019-00794-3
  • [AMZ20] Alexander Alldridge, Christopher Max and Martin R Zirnbauer “Bulk-boundary correspondence for disordered free-fermion topological phases” In Communications in Mathematical Physics 377.3 Springer, 2020, pp. 1761–1821
  • [ABJ20] Joachim Asch, Olivier Bourget and Alain Joye “On stable quantum currents” In Journal of Mathematical Physics 61.9 AIP Publishing LLC, 2020, pp. 092104
  • [BS20] Chris Bourne and Hermann Schulz-Baldes “On Z 2-indices for ground states of fermionic chains” In Reviews in mathematical physics 32.09 World Scientific, 2020, pp. 2050028
  • [Fon+20] Eli Fonseca, Jacob Shapiro, Ahmed Sheta, Angela Wang and Kohtaro Yamakawa “Two-Dimensional Time-Reversal-Invariant Topological Insulators via Fredholm Theory” In Mathematical Physics, Analysis and Geometry 23.3, 2020, pp. 29 DOI: 10.1007/s11040-020-09342-6
  • [Sha20] Jacob Shapiro “The topology of mobility-gapped insulators” In Letters in Mathematical Physics 110.10, 2020, pp. 2703–2723 DOI: 10.1007/s11005-020-01314-9
  • [BO21] Chris Bourne and Yoshiko Ogata “The classification of symmetry protected topological phases of one-dimensional fermion systems” In Forum of Mathematics, Sigma 9 Cambridge University Press, 2021, pp. e25 DOI: 10.1017/fms.2021.19
  • [Sha21] Jacob Shapiro “Incomplete Localization for Disordered Chiral Strips”, 2021 arXiv:2108.10978 [math-ph]
  • [Zir21] Martin R Zirnbauer “Particle–hole symmetries in condensed matter” In Journal of Mathematical Physics 62.2 AIP Publishing, 2021
  • [AT22] Joseph E. Avron and Ari M. Turner “Homotopy of periodic two by two matrices”, 2022 arXiv:2212.07529 [math-ph]
  • [Gei22] Tobias Geib “Topological classification of symmetric quantum walks. Discrete symmetry types and chiral symmetric protocols” Hannover: Institutionelles Repositorium der Leibniz Universität Hannover, 2022
  • [GMP22] David Gontier, Domenico Monaco and Solal Perrin-Roussel “Symmetric Fermi projections and Kitaev’s table: Topological phases of matter in low dimensions” 041902 In Journal of Mathematical Physics 63.4, 2022 DOI: 10.1063/5.0084326
  • [SW22] Jacob Shapiro and Michael I. Weinstein “Is the continuum SSH model topological?” 111901 In Journal of Mathematical Physics 63.11, 2022 DOI: 10.1063/5.0064037
  • [SW22a] Jacob Shapiro and Michael I. Weinstein “Tight-binding reduction and topological equivalence in strong magnetic fields” In Advances in Mathematics 403, 2022, pp. 108343 DOI: https://doi.org/10.1016/j.aim.2022.108343
  • [BSS23] Alex Bols, Jeffrey Schenker and Jacob Shapiro “Fredholm Homotopies for Strongly-Disordered 2D Insulators” In Communications in Mathematical Physics 397.3, 2023, pp. 1163–1190 DOI: 10.1007/s00220-022-04511-w