This paper proposes a classification of elliptic (pseudo-)differential Hamiltonians describing topological insulators and superconductors in Euclidean space by means of domain walls. Augmenting a given Hamiltonian by one or several domain walls results in confinement that naturally yields a Fredholm operator, whose index is taken as the topological charge of the system. The index is computed explicitly in terms of the symbol of the Hamiltonian by a Fedosov–Hörmander formula, which implements in Euclidean spaces an Atiyah–Singer index theorem. For Hamiltonians admitting an appropriate decomposition in a Clifford algebra, the index is given by the easily computable topological degree of a naturally associated map. A practically important property of topological insulators is the asymmetric transport observed along one-dimensional lines generated by the domain walls. This asymmetry is captured by the edge conductivity, a physical observable of the system. We prove that the edge conductivity is quantized and given by the index of a second Fredholm operator of the Toeplitz type. We also prove topological charge conservation by stating that the two aforementioned indices agree. This result generalizes to higher dimensions and higher-order topological insulators, the bulk-edge correspondence of two-dimensional materials. We apply this procedure to evaluate the topological charge of several classical examples of (standard and higher-order) topological insulators and superconductors in one, two, and three spatial dimensions.
I. INTRODUCTION
This paper considers topological insulators modeled by single particle Hamiltonians in the form of (pseudo-)differential operators (PDOs). See, e.g., Refs. 11, 43, 47, 48, 51, 56, and 57 for background and details on these materials and topological phases of matter. The partial differential systems we consider here naturally appear in a variety of contexts, which include heuristic descriptions of topological insulators and superconductors,11,56,57 macroscopic approximations of more accurate microscopic models such as, for instance, tight-binding11,43 or modulated periodic Schrödinger equations,33,34 as well as models of atmospheric fluid flows.24,54
Two-dimensional topological insulators enjoy the following striking property. When two insulators in different topological phases are brought together, the interface separating them becomes conducting, i.e., no longer insulating. Moreover, transport along the interface is asymmetric with an asymmetry quantized, and stable against perturbations. This guaranteed asymmetric transport in the presence of impurities is one of the main practical interests of such materials. It is also surprising since the non-trivial topology may be interpreted as an obstruction4,47 to Anderson localization, which states that transport is exponentially suppressed in the presence of random fluctuations.
Two main mechanisms of confinement lead to insulators. The first one is the magnetic confinement found in the integer quantum Hall effect, the first example of a topological insulator.1,2,10,11 Asymmetric transport along interfaces separating different magnetic confinements is analyzed for Schrödinger and Dirac equations in, e.g., Refs. 26 and 50. In such models, the magnetic confinement leads to flat bands of the essential spectrum (the Landau levels). The classification and many of the mathematical results we present below do not apply (potentially with false predictions) in the presence of such flat bands.
We are interested here in the second mechanism of confinement based on what we will refer to as mass terms. Mass terms typically take the form of one or more scalar functions. In two space dimensions, a scalar function m(x) may also act as a transition between the two insulating materials: the domains m > 0 and m < 0 are insulating, while the (vicinity of the) interface described as the 0-level set m−1(0) and separating the two bulk phases is conducting. We refer to such transition functions m(x) as domain walls.
It remains to define and compute topological invariants that characterize such an asymmetry. While such computations may be carried out explicitly for some models, typically by means of spectral flows,3,6,26,41,50 they remain notoriously difficult for more complex systems. A general principle called a bulk-edge correspondence relates the transport asymmetry along the interface to invariants associated with the bulk properties of the insulator in the regions ±m(x) > 0. The computation of such bulk invariants, typically by means of Chern numbers and winding numbers, is often much easier. Bulk-edge correspondences have been derived mathematically in a number of settings for discrete29,35,38,47,53 as well as continuous Hamiltonians;6,19,20,27,44,49 see also Refs. 30 and 56 for a bulk-boundary correspondence and a notion of topological charge conservation similar to the one we will describe in this paper. Directly related to this work is the correspondence established in Refs. 6 and 49, where the asymmetric transport is related to a bulk-difference invariant evaluated as a simple integral [the Fedosov–Hörmander formula in (10) below]. We remark that bulk phases for domains with m > 0 or with m < 0 constant may not be defined for natural differential operators such as the Dirac operator.3,6 In such settings, one may define a bulk-difference invariant6 that combines the properties of both domains m > 0 and m < 0. Heuristically, this indicates that it is easier to define phase transitions rather than absolute (bulk) phases.
This paper aims to generalize the above correspondence to arbitrary dimensions and for a large class of elliptic pseudo-differential Hamiltonians. The topological classification of each Hamiltonian is performed in two different ways. We start from a given Hamiltonian in d dimensions denoted by Hk, where 0 ≤ k ≤ d − 1 refers to the number of confined dimensions. Therefore, H0 with k = 0 may be a bulk metallic system (a gapless Hamiltonian such as a Dirac model for transport in graphene). In dimension d = 2, k = 1 may correspond to a Hamiltonian describing a topological transition between two insulators modeled by m(x1) > 0 and m(x1) < 0, respectively. We wish to test the topological properties of Hk by assessing its response to the addition of domain walls. When k ≤ d − 2, we first construct a Hamiltonian Hd−1 by appropriately adjoining d − k − 1 domain walls to it. The resulting Hamiltonian Hd−1 models a system that is confined in all but one dimension, say xd. We, therefore, expect asymmetric transport to occur along the corresponding dth dimension. We will introduce an edge conductivity σI(Hd−1) in (11) in Sec. III; this is a physical observable of the system associated with the asymmetric transport. We will show that 2πσI is equal to the index of a Fredholm operator T, which is a Toeplitz operator. Such Fredholm operators and the structures of Fredholm modules (or spectral triples) naturally appear in topological classifications.6,19,20,27,44,47,49 As in the two-dimensional setting, the computation of such an index is difficult in practice.
Following Refs. 6 and 49, we, therefore, test the topology of Hk by implementing a final domain wall m(xd) in the dth dimension and introduce an operator F = Hd−1 − im(xd). We will show that F is a (non-Hermitian) Fredholm operator. The topological charge of Hk is then defined as the Index F = dim Ker F − dim Ker F*. The main advantage of this classification is its computational simplicity: the index of F is given by an explicit integral of its symbol, the Fedosov–Hörmander formula (10) below. Moreover, the mapping from Hk to F is local in the sense that F = γ1 ⊗ Hk + γ2 for γ1 a constant matrix and γ2 an explicit multiplication operator in the physical variables. The evaluation of (10) is thus directly related to the symbol of a given Hamiltonian Hk of interest. For technical reasons that will allow us to apply the pseudo-differential calculus and index theory presented in Ref. 40, Chap. 19, we assume here that all (smooth) domain walls are unbounded. Since transport is exponentially suppressed away from the 0-level sets of domain walls, this is a reasonable assumption in many practical settings. See Refs. 6 and 49 for a theory in two dimensions involving bounded domain walls.
The above constructions describe two topological charges associated with Hk and given by the indices of the Fredholm operators F and T. As a generalization of the bulk-edge correspondence, the main result of this paper is that the topological charge is conserved through a series of deformations leading to an equality of the two indices. This shows that the physically relevant asymmetric transport characterized by the edge conductivity and the Fredholm operator T may be estimated by the much simpler integral (10) associated with F.
As a final theoretical result, we show that for specific Hamiltonians, the integral (10) admits an explicit expression given as the topological degree of a natural map associated with Hk. Such degrees, which naturally take integral values, are often quite easy to compute in problems of interest. We will see that the construction of Hd−1 and F from Hk in arbitrary dimensions requires a structure of Clifford algebras in order to define appropriately orthogonal domain walls. When Hk itself admits a natural decomposition in a Clifford algebra, then we will associate with Hk a map hk whose topological degree is straightforward to compute for several practical Hamiltonians.
The classification based on domain walls may be compared with other classification mechanisms. Topological phases of matter are heuristically characterized by non-trivial topologies of Hamiltonians in dual, Fourier, variables.11,47,56 This non-trivial topology may be tested in several ways. Linear response theory in two dimensions tests a domain by applying, e.g., a linear electric field in one direction and assessing the resulting transport in the transverse direction leading to the notion of Hall conductivity.2,11,36 While physically different, adding to a metallic operator H0 in two dimensions a domain wall in a given direction and testing asymmetric transport in the transverse direction play a similar classifying role. An advantage of the classification based on domain walls is its natural generalization to arbitrary dimensions and the explicit Fedosov–Hörmander formula satisfied by the index. Note that the latter is also used in a different context (and with a different algebra) of operators in Euclidean space by Refs. 18 and 21 to test the topology of a physical potential with appropriate behavior at infinity in the physical variables using a Dirac operator.
The rest of the paper is structured as follows: the definition and construction of the operators Hk, Hd−1, F, and T require an unavoidable amount of pseudo-differential and functional calculus. The conservation of topological charge, as in the index theory developed in Ref. 40, Chap. 19, requires us to consider semiclassical transformations and an associated semiclassical calculus. The notation and required results, which may be found in more detail in Refs. 15, 25, and 40, are collected in the Appendix.
The functional setting and construction of the operator Hd−1 and F using confinement by domain walls are presented in Sec. II. The main result of this section is the index formula in Ref. 40, Chap. 19 recalled in Theorem 2.3. The definition of the edge conductivity σI associated with Hd−1 is given in (11) in Sec. III. We define a corresponding Fredholm operator in Toeplitz form and show in Theorem 3.2 that 2πσI = Index T. Several lemmas in that section also show the stability of 2πσI = Index T against a number of continuous deformations of the Hamiltonian Hk.
The conservation of topological charge Index F = Index T, the main theoretical result of this paper, is proved in Theorem 4.1 of Sec. IV. Finally, the computation of the integral (10) by means of the topological degree of the map hk associated with Hamiltonians Hk in appropriate Clifford algebra form is carried out in Theorem 5.1 in Sec. V.
Section VI is devoted to several applications of the theoretical results. It details the computation of the topological charge of generalizations of Dirac systems of equations in Clifford algebra form, possibly including magnetic contributions, as well as some Hamiltonians not in Clifford algebra form. An application to higher-order insulators in three spatial dimensions shows, in particular, how two domain walls may be defined to construct a coaxial cable with an arbitrary number of topologically protected propagating modes. These models find applications in, e.g., graphene- and bilayer graphene-based topological insulators, topological superconductors, and topological atmospheric flows.
II. FREDHOLM OPERATOR AND TOPOLOGICAL CHARGE
This section introduces classes of elliptic pseudo-differential Hamiltonians, classifies them by means of domain wall confinements, and defines their topological charge as the index of a Fredholm operator naturally associated with each Hamiltonian. The advantage of this classification is the explicit integral (10) that computes the topological charge. This Fedosov–Hörmander formula, presented in Theorem 2.3 below, implements in Euclidean geometry an Atiyah–Singer theory.12
A. Classification of Dirac operators in low dimensions
To illustrate how the topology of Hamiltonians is tested by domain walls, we present the constructions for Dirac operators, which are generic models for band crossings11,27,32–34,51 and arguably the simplest models for topological phases of matter.
Consider first a one-dimensional material and asymmetric transport modeled by the Hamiltonian H0 = Dx with Dx = −i∂x, which may be seen as an unbounded self-adjoint operator on with domain . This operator admits purely absolutely continuous spectrum and displays asymmetric transport along the x axis in the sense that solutions of (Dt + Dx)u = 0 with initial condition u(0, x) = u0(t) are given by u(t, x) = u0(x − t).
To generate confinement in the vicinity of x = 0, we introduce the domain wall m(x) = x and the operator with an annihilation operator. The operator F is now a Fredholm operator from its domain of definition to . Moreover, we verify that Index F = dim Ker F − dim Ker F* = 1 with kernel of F spanned by the eigenfunction . We define Index F as the topological charge of H0.
Consider next the two-dimensional version of the above example, where H0 = D1σ1 + D2σ2 with Dj = −i∂j for j = 1, 2 and σ1,2,3 are the standard Pauli matrices. The operator H0 appears as a generic low-energy description of energy band crossings and is ubiquitous in works on topological insulators. We classify H0 by augmenting it with a domain wall along one direction and assessing the resulting asymmetric transport in the transverse direction. We implement a domain wall m1(x1) = x1 along the first variable by introducing H1 = H0 + x1σ3. This models includes insulating regions x1 > 0 and x1 < 0 while transport remains possible in the vicinity of the interface x1 = 0.
To obtain confinement in the second variable, we introduce the second domain wall m2(x2) = x2. Associated to H0 and H1 is then the operator F = H1 − ix2. This is again a Fredholm operator from its domain of definition to and we verify that Index F = 1, which defines the topological charge associated with H0 (and H1). The kernel of F is spanned by the spinor , while the kernel of F* is trivial.
This construction generalizes to higher dimensions in a straightforward way, except for the fact that the construction of the domain walls requires additional degrees of freedom as dimension increases. Note that in dimension d = 1, we have with c1 = −1 here. In dimension d = 2, we observe that with I2 the 2 × 2 identity matrix. The latter is obtained because (σ1, σ2, σ3, i) satisfy appropriate orthogonality properties. In particular, σ1,2,3 satisfy σiσj + σjσi = 2δijI2. Ensuring this property in higher spatial dimensions requires enlarging the size of the spinors so they generate an appropriate representation of the Clifford algebra (Ref. 12, Chap. 17).
Consider in the Weyl Hamiltonian H0 = D1σ1 + D2σ2 + D3σ3. As an operator acting on spinors in , the latter operator is stable against gap opening by domain walls.11 We, therefore, introduce the operator H1 = σ1 ⊗ H0 + σ2 ⊗ I2x1 with a domain wall m1(x1) = x1 in the first direction but now acting on spinors in . The operator H1 thus admits surface states concentrated in the vicinity of x1 = 0, as does the operator H0 in the two-dimensional setting. Its topology is then characterized by asymmetric transport in the third dimension after a second domain wall in the x2 direction is introduced: H2 = H1 + σ3 ⊗ I2x2.
Confinement in the last variable is imposed by the domain wall m3(x3) = x3. We introduce the operator
This is a Fredholm operator [from its domain of definition to ]. We verify (and will show in greater generality in Sec. VI) that the topological charge of H0 is Index F = −1, with a sign change here reflecting the fact that indices depend on the orientation of the Clifford matrices used to construct the operators Hj as well as the orientation of the domain walls. The kernel of F* has for eigenfunction the spinor . The topological charge of H0, of H1, and of H2 is defined as Index F.
Note that in the above three-dimensional model, H2 models higher-order topological insulators with transport occurring along a hinge, or co-dimension two manifold.52 The classification of H2 by domain walls is arguably simpler than one based on bulk phases that may be hard to identify.
B. Pseudo-differential elliptic operators
We now define the classes of pseudo-differential operators we wish to classify. Such classes naturally generalize the examples of Dirac operators seen in the preceding paragraphs.
Consider a spatial dimension d ≥ 1 and operators defined on functions of the Euclidean space . We denote by the dual (Fourier) variable and the phase space variable. The algebras of pseudo-differential operators (PDOs) we consider are written in Weyl quantization as
for a(X), a matrix-valued symbol in , the space of complex n × n matrices, and ψ(x) a spinor with values in . The value of n is problem-dependent. For the Fredholm operator (1), we verify that F = Op a for a(X) = σ1 ⊗ (ξ1σ1 + ξ2σ2 + ξ2σ3) + x1σ2 ⊗ I2 + x2σ3 ⊗ I2 − ix3I4 with thus n = 4. Relevant notation and results on pseudo-differential operators and functional calculus are collected in the Appendix.
The starting point is an operator Hk = Op ak with ak a given matrix-valued symbol in interpreted as confining the first k variables. Our first aim is to construct the operators H ≔ Hd−1 = Op ad−1 and F = Op a with F a Fredholm operator. This requires introducing the following notation and functional setting.
We decompose the spatial variables with and . We use the notation and and define the weights
For a given spinor dimension n = nk with 0 ≤ k ≤ d, and an order m > 0, we denote by the class of symbols ak such that for each d-dimensional multi-indices α and β, there is a constant Cα,β such that for each component b of , we have
Here, . We define the space of symbols as but acting on spinors of (lower) dimension nd−1 instead of nd. Here m is the order of the operator.
For the two-dimensional Dirac operator, we find m = 1, H0 = Op a0 for a0 = ξ1σ1 + ξ2σ2, while H1 = Op a1 for a1 = a0 + x1σ3 and F = Op a for a = a1 − ix2. For n0 = n1 = 2, we observe that for j = 0, 1, while .
We impose a number of assumptions on Hk and ak. The first one is Hermitian symmetry . A second symmetry is the chiral symmetry (6) below when d + k is even. This symmetry ensures the (potential) topological non-triviality of the Hamiltonian. Finally, we assume all our Hamiltonians to be elliptic, a necessary assumption in any index theory.12 Ellipticity is defined as the constraint
In other words, all eigenvalues of the Hermitian matrix ak(X) are bounded away from 0 by at least for X outside of a compact set in . Since , all (positive and negative) eigenvalues of ak are of order away from a compact set.
We denote by the elliptic symbols in and the elliptic symbols in .
Associated to the spaces of symbols and are Hilbert spaces and defined in (A5) in the Appendix. These spaces are constructed so that for , we obtain that Hk = Op ak is an unbounded self-adjoint operator on with domain of definition , while for , we obtain that F = Op a is an unbounded operator with domain of definition . For any , the space is defined explicitly as .
For the Dirac operators in dimension d = 3, we have, for instance, the Hilbert space with n0 = 2 and n1 = n2 = n3 = 4.
C. Classification by domain walls
We start from a given elliptic (self-adjoint) operator Hk = Op ak for with m > 0. By ellipticity assumption, Hk is an unbounded self-adjoint operator with domain and acts of spinors in . The ellipticity of Hk and the construction of the weight imply that the first k variables parameterized by are confined in the sense that each singular value of ak is large for large.
To obtain a non-trivial topological classification, the operator Hk needs to satisfy a chiral symmetry when d + k is even (complex class AIII47). When d + k is odd, no symmetry is imposed beyond the Hermitian structure.
Assume first that d + k is even with k ≤ d − 2. Recall that σ1,2,3 are the Pauli matrices, a set of Hermitian 2 × 2 matrices such that σiσj + σjσi = 2δij and σ1σ2 = iσ3. Using the notation , the chiral symmetry takes in a suitable basis the following form:
We next introduce the domain walls
They are constructed to have the same asymptotic homogeneity of order m as the Hamiltonian Hk. We then define the new spinor dimension nk+1 = nk and the augmented Hamiltonian
This implements a domain wall in the variable xk+1.
Assume now d + k is odd. We define the new spinor dimension nk+1 = 2nk and the augmented Hamiltonian,
The operator Hk+1 satisfies the chiral symmetry of the form (6), as requested since d + k + 1 is now even.
We denote by ak+1 the symbol of Hk+1 = Op ak+1 and observe that ak+1 = ak + mk+1σ3 ⊗ I when d + k is even and ak+1 = σ1 ⊗ ak + mk+1σ2 ⊗ I when d + k is odd.
The procedure is iterated until Hd has been constructed. Note that with dimension of the spinor space on which the matrices act that doubles every time k is raised to k + 2. Since 2d is even, Hd = σ− ⊗ F* + σ+ ⊗ F for an operator F = Fd = Hd−1 − imd =: Op a, or equivalently, a = ad−1 − imd.
For 0 < l ≤ d − k, the intermediate Hamiltonians all have the form
where for some integer p = p(l, k) and for some matrices γj such that {γi, γj} ≔ γiγj + γjγi = 0 for all i ≠ j in {0, …, l}, we have
We now show that all operators Hl are elliptic and that Hd and F are Fredholm operators.
Let and Hk = Op ak satisfying the chiral symmetry (6) when d + k is even. Then for all k ≤ j ≤ d and .
Let ak+l be the symbol of Hk+l = Op ak+l for 0 < l ≤ d − k. By construction and commutativity {γi, γj} = 0, we obtain that with I identity matrices with appropriate dimensions. This shows that ak+l satisfies the ellipticity condition (5) for the weight wk+l(X). The decay properties for derivatives of ak+l in (4) with k replaced by k + l follow from the corresponding properties for ak. That comes from the corresponding result for ad and the construction of F.
Let be an elliptic self-adjoint operator, which by construction, maps to .45
The operators Hd and F are Fredholm operators from to and to , respectively. Equivalently, Λ−mHd and Λ−mF are Fredholm operators on and , respectively.
The construction of Hl implements l − k domain walls to test the topology of the operator Hk. When l = d, the operator Hd has d confined variables and is, as we saw, a Fredholm operator, i.e., an operator that admits left and right inverse modulo compact operators (Ref. 40, Chap. 19). The operator Hd is self-adjoint, and so its index vanishes. However, it satisfies the chiral symmetry (6), and the corresponding operator F = Hd−1 − imd is also Fredholm. Its index may not vanish, and we define the topological charge56 of Hk as Index F.
D. Topological charge and integral formulation
We next apply Ref. 40, Theorem 19.3.1′ (see also Ref. 31) to the operator Λ−mF to obtain that the index of F, which equals that of Λ−mF since Λ−m has trivial index, is given by the following Fedosov–Hörmander formula:
(Ref. 40, Theorem 19.3.1′). Comes from the approximation of symbols in S(M, gs) by symbols in S(M, gi) and the index theorem (Ref. 40, Theorem 19.3.1) proving (10) for symbols in S(M, gi). The approximation is described in Ref. 40, Lemma 19.3.3. See the Appendix for notation on the metrics gi and gs. We use a similar approximation in Lemma 3.3 below to prove topological charge conservation in Theorem 4.1.
(Ref. 40, Theorem 19.3.1′). Applies to Λ−mF and not F directly. However, the index of Λ−tmF is independent of t ∈ [0, 1] since the index of Λ−tm is trivial and the index of a product is the sum of the indices (Ref. 40, Chap. 19). The corresponding symbols at of Λ−tmF are uniformly invertible for |X| ≥ R for R sufficiently large and smooth in t. The integral (10) is, therefore, the same for a = a0 and a = a1. This may be proved as follows:
Let M and N be smooth closed manifolds. Here, and . For t ∈ [0, 1], let it: M → M × [0, 1] defined by it(x) = (x, t). Let a0 and a1 be homotopic smooth maps from M to N, and let a: M × [0, 1] → N be the smooth homotopy map such that a0 = a◦i0 and a1 = a◦i1.
III. PHYSICAL OBSERVABLE AND TOEPLITZ OPERATOR
While systematic and explicit, the classification of Hk in Sec. II based on the index of F is a priori unrelated to any physical observable. We now present a second topological classification based on a physical observable that characterizes the asymmetric transport of the operator Hd−1 introduced in Sec. II C.
The operator Hd−1 confines in d − 1 directions while allowing transport in the remaining dimension parameterized by xd. The following edge conductivity quantifies asymmetric transport in that direction. Let H ≔ Hd−1 and a smooth non-decreasing switch function and a smooth spatial switch function.
Here and below, a function is called a switch function if f is bounded measurable and there are xL and xR in such that f(x) = 0 for x < xL and f(x) = 1 for x > xR. We denote by the subset of smooth switch functions.
We define the edge conductivity
Here, φ′(H) defined by functional calculus with φ′(h) ≥ 0 being a spectral density (integrating to 1). We assume that i[H, P]φ′(H) is a trace-class operator (i.e., a compact operator with summable singular values). The commutator is defined as [A, B] = AB − BA. The edge conductivity σI has been used to model edge transport and derive a bulk-edge correspondence for two dimensional materials in a number of contexts (see, e.g., Refs. 6, 27, 29, 36, 38, 47, and 49).
It may be given the following interpretation. Let ψ(t) = e−itHψ be a solution of the Schrödinger equation i∂tψ(t) = Hψ(t) with initial condition ψ, and let P be a Heaviside function defined as P(xd) = 1 for while P(xd) = 0 for for some . Then is interpreted as the mass of ψ(t) in the half space on the right of the hyperplane . Its derivative thus describes current crossing the hyperplane . We formally replace the density ψ(t)ψ(t)* by the spectral density φ′(H). This heuristically gives the interpretation of σI as the rate of signal propagating from the left to the right of the hyperplane per unit time for a density of states in the system given by the spectral density φ′(H).
The main objective of this section is to prove that for , then i[Hd−1, P]φ′(Hd−1) is indeed a trace-class operator so that σI(Hd−1) is well-defined. We next relate the edge conductivity to the index of the Toeplitz operator for an orthogonal projector in and U(H) = e2πiφ(H). In particular, we show that so that σI is indeed quantized. Finally, we prove that σI and the index of T are stable against a number of continuous transformations of the symbol ak.
A. Trace-class property and Toeplitz operator
Let so that by Lemma 2.1, while . We denote by H = Hd−1 = Op ad−1. Let U(H) = ei2πφ(H) with while W(H) = U(H) − I.
We use the following notation for classes of symbols. These classes are also recalled in the Appendix. For M = M(X), a weight function, we denote by S0(M) the space of symbols with components b(X) such that
for each multi-index α. Therefore, is a larger class of symbols than defined in (4). The pseudo-differential calculus provides the following results, as recalled in the Appendix:
If a ∈ S0(M1) and b ∈ S0(M2), then Op a Op b = Op c with c ∈ S0(M1M2) constructed explicitly in (A9) (with h = 1).
If and , then .
If a ∈ S(M) with , then Op a is trace-class and (A11) holds; namely, the trace of Op a is given by (2π)−d times the integral of the trace of a over , or equivalently by the integral of the trace of their Schwartz kernel along the diagonal; see (A7) (with h = 1).
This allows us to obtain the following results:
Let and . Let and . Then [P, ϕ(H)] and Hp[P, Hq]ϕ(H) are trace-class operators with symbols in . When is an orthogonal projector, then is a Fredholm operator on with index given by . All the above operator traces may be computed by integrating the Schwartz kernel of the operator along the diagonal.
By composition calculus (i), for any A = Op a with a ∈ S0(M), then the decomposition for a smooth function equal to 1 for x > 1 and 0 for x < −1 shows that [A, P] has symbol in (i.e., in for each ). By assumption on ak and using the functional calculus result (ii) (see Lemma A.2), we obtain for that , which is larger that (since m > 0). To simplify notation, we use the same notation for S0(M) and for any n. Therefore, by composition calculus, [ϕ(H), P] and [H, P]ϕ(H) as well as Hp[Hq, P]ϕ(H) for all have symbols in S0(⟨X⟩−∞). We use this with ϕ = φ′ and ϕ = W. With additional effort, we verify that all symbols are in , although this is not necessary for the rest of the proof and so we leave the details to the reader.
Using (iii), we deduce that [P, ϕ(H)] and Hp[P, Hq]ϕ(H) are trace-class operators with traces given as the integral of their Schwartz kernel along the diagonal. Applying the latter directly yields that Tr[ϕ(H), P] = 0, for instance, since the Schwartz kernel of [ϕ(H), P] vanishes along the diagonal. In particular, [P, W(H)] and [H, P]φ′(H) are trace-class operators.
That is a Fredholm operator with index given by for and projectors is a non-trivial consequence of the trace-class nature of and the Fedosov formula (see, e.g., Ref. 2, Proposition 2.4).
We now relate the Fredholm operator T and the calculation of its index as a trace with the line conductivity σI = σI(H) defined in (11). We obtain the following result:
Under the assumptions of Lemma 3.1, we have .
B. Stability results
We next derive several results showing that the index of the Toeplitz operator is stable against a number of (continuous) deformations. We first need a technical result showing that the index can be computed by approximating the symbol of Hk by an isotropic symbol. All results so far have been obtained for symbols satisfying (4). The space of isotropic elliptic symbols is defined by a similar constraint where ⟨x⟩|α|⟨ξ⟩|β| is replaced by ⟨X⟩|α|+|β|. The class of elliptic isotropic symbols is thus smaller than . It is also invariant under permutation of the variables X, which is not the case for . The invariance under such permutations will be needed in the proof of conservation of the topological charge in Sec. IV. We thus state the following approximate result:
Let with . Then there is a sequence of operators Hɛ for 0 ≤ ɛ ≤ 1 with symbol in for all ɛ > 0 and such that the corresponding is continuous in the uniform sense and T0 = T. Therefore, Index Tɛ is defined as being independent of ɛ and equal to Index T. Moreover, the symbols aɛ are chosen so that for any compact domain in X = (x, ξ), aɛ = ad−1 on that domain for ɛ sufficiently small.
We now mimic the proof of Ref. 40, Theorem 19.3.1′ extending the index theorem (10) from the isotropic metric gi to the metric gs. For aɛ(X) = ad−1(v(ɛ|X|)x, v(ɛ|X|)ξ), we find that uniformly in ɛ and for ɛ > 0. Let Hɛ = Op aɛ and . By uniformity of in ɛ and uniformity of bounds in Lemma A.2 (see Remark A.3), we obtain that W(Hɛ) has symbol in uniform in ɛ.
We next observe that and are uniformly compact in ɛ and even uniformly trace-class from the results of Lemma 3.1. We then apply Ref. 40, Theorem 19.1.10 to obtain that the indices of and , and hence that of Tɛ, are independent of 0 < ɛ ≤ 1. In the limit ɛ → 0, this is the index of T. Therefore, Index T = Index Tɛ for ɛ > 0 but now for a symbol .
The above result shows that we can replace the symbol in Hd−1 with that obtained at ɛ > 0. We also observe that (10) is independent of ɛ for ɛ small. We may, therefore, assume that in the computation of Index T. The main advantage of the more constraining metric gi is that the corresponding symbol classes are now invariant under suitable rotations and permutations of the phase space variables X. The following result is then used. Let Y = (x1, …, xd−1, ξ1, …, ξd−1).
Let g = gi. Let [0, 1] ∋ t → Lt be a continuous family of linear invertible transformations in in the Y variables leaving the variables (xd, ξd) fixed. Let . Then . Let Tt be the corresponding Toeplitz operator. Then Tt is Fredholm with index independent of t ∈ [0, 1].
Higher-order derivatives are bounded in the same way, allowing us to obtain that is bounded by a constant times |t − s|. Therefore, U(Ht) − U(Hs) is small in the uniform sense for small (t − s), so that the index of Tt is continuous in t and hence independent of t ∈ [0, 1].
This result states, in particular, that for ad−1 ∈ ESd−1(gi), then the index of T is independent of any rescaling Yj → λYj for λ > 0 (leaving all other variables fixed) as well as any rotation in the phase space variables mapping (Yj, Yk) to (Yk, − Yj) (note the sign change to preserve orientation). We will use the above lemma only for such transformations (dilations and permutations).
We could show similarly that σI is independent of changes in φ although this property will automatically come from (14) proved in Sec. IV. However, we need the following straightforward result:
Let g ∈ {gs, gi}. For t ∈ [0, 1], let be a continuous path of elliptic symbols. Then the indices of the corresponding Fredholm operators Ft = Op (at − imd) and are independent of t ∈ [0, 1].
By assumption and construction, Λ−mFt and Tt are continuous in t as operators from to itself. Their indices are, therefore, constant in t.
We apply the preceding lemma to and while at = ta1 + (1 − t)a0. The path of symbols belongs to so that the indices of the respective operators are defined with clear continuity in t. This allows us to replace the xd-dependent elliptic symbol ad−1 with an xd-independent one, which is used below in the proof of topological charge conservation.
IV. TOPOLOGICAL CHARGE CONSERVATION
We recall that Hk = Op ak for an elliptic symbol with 0 ≤ k ≤ d − 1 and m > 0. We constructed in Sec. II, with operator Hd−1 = Op ad−1 confined in all but the last variable and a Fredholm operator F = Op a confined in all variables. Associated to Hd−1 is the edge conductivity σI defined in (11). We associated two topological charges to F and σI in Theorems 2.3 and 3.2, respectively. This section shows that the two classifications are in fact equivalent and that we have the following conservation of the topological charge:
As a first step, we continuously deform to using Lemma 3.5 and the paragraph that follows it. Note that all terms in (14) are stable under this change of symbols (see Remark 2.5). We next use the approximation of a symbol in by using Lemma 3.3. Note that again, all terms in (14) are stable under this change of symbols since both symbols agree on the support of .
To simplify the presentation, we change notation to and to (ζ, ξ) = ξ with the new and . Therefore, the symbol ad−1 = ad−1(y, ζ, ξ).
We finally continuously deform a(X) to ad−1(y, ζ, ξ) − ix. This does not change the definition of σI. Using the homotopy result of Remark 2.5, this does not change the integral over either by continuously deforming the mass term to xd ≡ x (using for t ∈ [0, 1]) on the ball of radius R while preserving the continuity of a−1 on .
For the same reason, we may replace a by ad−1(y, ζ, ξ) + α − ix for any by continuity in α and for a fixed radius R for α in a compact domain.
We now introduce the semiclassical parameter 0 < h ≤ 1 and the operator with the symbol ad−1(y, hζ, ξ). Using the semiclassical notation recalled in the Appendix (in the phase-space variable Y), we thus observe that . Using Lemma A.1, we know that is a PDO with the semiclassical symbol rz(y, ζ, ξ; h). Note that the latter term has a complicated dependence on h. We know from Lemma 3.4 that σI(Hh) is independent of 0 < h ≤ 1. We may, therefore, compute it in the limit h → 0.
The integral on is also independent of λ in a compact interval since λ → σz is continuous as an application of the result in Remark 2.5. Since and σα+iω(ξ, Y) = −a(ω, ξ, Y), the integrals in (14) and (22) agree modulo a sign. Upon inspection, we observe that the integral in (22) has been computed for the orientation dxd ∧ dξd ∧ dx1 ∧ dξ1…dxd−1 ∧ dξd−1 > 0 with ω ≡ xd the first variable defining the surface S. The latter orientation is (−1)ddξ1 ∧ dx1 ∧ ⋯∧ dξd ∧ dxd. With the latter choice of orientation, we obtain the topological charge conservation between the topological charge given by the index of F and the transport asymmetry given by the conductivity 2πσI stated in (14).
In dimension d = 2 and in the setting of bounded domain walls, the Fedosov–Hörmander formula may be interpreted as a difference of bulk quantities since the integral over the sphere may be deformed into the integral over two hyperplanes in the bulk phases where x1 = ±R constant. This also uses as in the above derivation that and the Stokes theorem (see Ref. 6). The topological charge conservation of Theorem 4.1 thus generalizes the two-dimensional bulk-interface correspondence to an arbitrary space dimension.
V. GENERALIZED DIRAC OPERATORS AND DEGREE THEORY
We assume in this section that has the following form:
where hk(X) is a (d + k)-dimensional vector field on and Γk is a collection of matrices in a representation of the Clifford algebra ; see, e.g., Ref. 47 for details on Clifford algebras and their central role in the analysis of topological insulators. The objective of this section is to show that the index of the Fredholm operator F associated with Hk = Op ak may be computed as a topological degree associated with the map hk. This simplifies the estimation of the integral (10).
For 0 ≤ k ≤ d, let and . The matrices for 1 ≤ j ≤ d + k are constructed to satisfy the commutation relations
These properties imply that is proportional to identity. The matrices Γk may be defined explicitly as follows: the matrices at level κ are constructed starting from the standard Pauli matrices and then iteratively as
The last matrix plays the role of the chiral symmetry matrix in even dimension d + k = 2κk. The construction of the augmented Hamiltonians Hj for k < j ≤ d in Sec. II mimics the construction of the above matrices. When d + k is even, the chiral symmetry is implemented as
For ak = hk · Γk, we denote by hj for k ≤ j ≤ d the vector fields of dimension d + j such that the augmented Hamiltonians constructed in Sec. II satisfy Hj = Op aj with, as we verify, aj = hj · Γj.
Dirac operators are the prototypical example of operators in the form (23). In two dimensions, we have explicitly Γ0 = (σ1, σ2) while h0(X) = (ξ1, ξ2) and Γ1 = (σ1, σ2, σ3) while h1(X) = (ξ1, ξ2, x1). In dimension d = 3, we have Γ0 = (σ1, σ2, σ3) while h0(X) = (ξ1, ξ2, ξ3), next Γ1 = (σ1 ⊗ σ1, σ1 ⊗ σ2, σ1 ⊗ σ3, σ2 ⊗ I2) while h1(X) = (ξ1, ξ2, ξ3, x1), and finally Γ2 = (σ1 ⊗ σ1, σ1 ⊗ σ2, σ1 ⊗ σ3, σ2 ⊗ I2, σ3 ⊗ I2) while h2(X) = (ξ1, ξ2, ξ3, x1, x2). When d = 3, then κ0 = 1 while κ1 = κ2 = 2 for a maximum of matrices satisfying (24) equal to 2κ2 + 1 = 5. Several other examples will be presented in Sec. VI.
A. Topological charge computation
For elliptic operators that admit the Clifford representation (23), the explicit computation of the index in (10) significantly simplifies as does the computation of the degree of the map hk.
We recall the definition of the degree of a map following Ref. 46, Chaps. 1.3 and 1.4; see also Ref. 28, Chaps. 13 and 14. Let C be an open set in with compact closure . Let be a sufficiently smooth map such that |h(ζ)| > 0 for ζ ∈ ∂C. There are regular values y0 of h arbitrarily close to 0 by Sard’s theorem that allow us to define the degree of h as
The above sum ranges over a finite set and is independent of the regular value y0 in an open vicinity of 0.
The definition of the index of a map from a manifold M to another manifold N depends on the chosen orientation on M. We consider two natural orientations in the context of topological insulators. Let the ball of radius R given by {|X| ≤ R}. By ellipticity assumption, |hd| > 0 on ∂Bd for R large enough. We now define degrees for hd with two possible orientations,
We observe that
The degree is naturally related to Index F, while the degree deg is more naturally related to that of hk as we now describe.
Using Lemma 3.5, we obtain that the index is unchanged if is replaced by in the definition of the symbol. We may, therefore, interpret hk as a map from to such that, thanks to the ellipticity constraint, |hk| ≥ h0 > 0 for . Let . We define
The orientation of Bk is inherited from that of Bd as the subset . With these definitions, we obtain the main result of this section.
In other words, Index F = deg(hk) when d = 1, 2 mod 4 and Index F = −deg(hk) when d = 3, 4 mod 4.
The rest of this section is devoted to the proof of the theorem. Its main steps are as follows: (i) write (10) in terms of σd; (ii) next in terms of hd; (iii) identify (10) with the degree of hd on the sphere ; (iv) identify it with the degree ; (v) decompose with the augmentation map. The degree of hd is then the product of the other two degrees. Now the degree of is one and this gives the result.
We recall that is the set of permutations of {1, …, n}. The proof of the lemma directly comes from the construction of the Clifford matrices in (25) (and their orientation) and generalizes that trσ3σ1σ2 = 2i. The above three lemmas show that the index of F is related to an appropriate integral of hd.
Let f be the Gauss map associated with hd and given by f(X) = |hd(X)|−1hd(X) for . Then we recognize in the integration of the right-hand side of (31) over the degree of f (Ref. 28, Corollary 14.21). Moreover, the degree of the Gauss map f is given in Ref. 28, Theorem 14.4.4 precisely by the sum in (26) and so equals when 0 is a regular value of hd. When 0 is not a regular value, we apply the result of hd − y0 for y0 small with the result independent of y0; see also Ref. 46, Remark 1.5.10. With the chosen orientation to define and Theorem 2.3, we thus obtain the result of the lemma.
We have deg(hd) = deg(hk).
We now choose h1 = hk and h2 the vector so that hd = (hk, h2) with n1 = d + k and n2 = d − k where is oriented using dξ1…dξddx1…dxd > 0 and the subspaces (for hk) and (for h2) with the induced orientation. We observe that the degree of equals 1 since the only point in and the Jacobian is identity there with the above orientation. Using (32) and the definitions (27) and (29) proves the result.□
The above lemmas together with the change of orientation relation (28) conclude the Proof of Theorem 5.1.
VI. APPLICATIONS
The classification presented in Sec. II applies to Hamiltonians that are (a) continuous (with an open “Brillouin” zone ), (b) defined on Euclidean space , and (c) appropriately elliptic with a symbol that tends to infinity at infinity in the variables (x′, ξ). Besides these constraints, the Hamiltonians are general when d + k is odd and to Hamiltonians with a chiral symmetry when d + k is even. The Fedosov–Hörmander formula (10) shows that the index is controlled by the symbol a of F, and hence that of Hk, restricted to any sphere with a sufficiently large radius R. This implies that the topological charge is independent of the symbol a in the complement of that ball. The main assumptions to apply (10) are that: (i) the symbol a of F is uniformly invertible for |X| ≥ R for some R > 0, in which case: (ii) the topological charge solely depends on a restricted to the sphere |X| = R.
The theory of Sec. II applies only to operators whose symbols satisfy the ellipticity constraint (5), which combined with the growth condition (4) implies that the symbol a grows to infinity as |X| → ∞ with the same homogeneity in all phase-space variables. This should be contrasted to the two-dimensional results in Refs. 6 and 49, where the domain wall m(x1) is assumed to be bounded and constant away from a compact domain.
For any symbol ak such that (i) and (ii) hold, we allow for the following modifications of the symbol ak in order to apply the theory of Sec. II. Let ɛ > 0 and r → ⟨r⟩ɛ a smooth non-decreasing function from to such that
We use the same notation for the smooth function . This function has the same leading asymptotic behavior as ⟨ɛy⟩ for |y| → ∞. We consider the above regularization for y being one or several of the variables in X. Such modifications of ak preserve (i)-(ii) and allow us to satisfy (5) as well as (4) so that the theory of Sec. II applies.
Consider, for instance, the regularized “Dirac” operator H1 = D1σ1 + D2σ2 + (μ − ηD · D)σ3 with here the (positive) Laplacian and . The definition of a bulk invariant is ambiguous when η = 0, while it yields a Chern number when η ≠ 0.3,11 As mentioned a number of times already in this paper, we do not consider bulk invariants but rather topological charges and interface invariants, which in two space dimensions may be related to bulk-difference (rather than bulk) invariants.6 To define a topology in the class of symbols analyzed in this paper and satisfy (5), we modify the above Hamiltonian as
where we assume that μ(x1) equals x1, say, outside of a compact set in . We will verify below that the topological charge of H1 equals 1 and is independent of the regularization terms ɛ and η as expected since η affects the bulk invariants but not the bulk-difference invariant.3 Note that (i) and (ii) now hold with m = 1. Alternatively, we could introduce satisfying (i) and (ii) with now m = 2.
We next consider several examples of topological insulators and superconductors in dimensions d = 1, 2, and 311,48,51,52,56 where the theories of both Secs. II and V apply. We refer to Ref. 9 for an application to Floquet topological insulators, where a variation on Theorem 4.1 is used to compute invariants for operators that are not in the form (23).
While the theory leading to Theorems 4.1 and 5.1 applies to a large class of practical settings, as the rest of this section illustrates, it does not apply to situations where the ellipticity conditions are not met. This is the case for confinements generating flat bands, for instance in Schrödinger or Dirac equations with magnetic fields, which may be analyzed by other techniques.26,50 This is also the case for the 3 × 3 Hamiltonian (39) describing fluid waves before regularization. See (39) and the following paragraphs for a regularized version.
A. Dirac operator
The first example is the Dirac operator with H0 = Op a0 for in dimension d, where
and Γ0 are Clifford matrices acting on spinors in with . These generalize the cases d = 1, 2, and 3 considered in the introduction. We then observe from (26) that deg h0 = 1 since and ∇h0(0) = Id, and that the topological charge of H0 is given by , i.e., Index F = 1 in dimensions 1, 2 mod 4 and Index F = −1 in dimensions 3, 4 mod 4.
The topological charge is given by the degree of h0 or by that of hd = (ξ1, …, ξd, x1, …, xd) since the deg h0 = deg hd = 1.
If A is a non-singular (constant) matrix in and we consider instead the operator H0 = Op (Ah0)(ξ) · Γ0, then we find that Index F = sign det A in dimensions 1, 2 mod 4 and Index F = −sign det A in dimensions 3, 4 mod 4.
The topological charge is also stable against large classes of smooth perturbations of arbitrary amplitude, so long as the perturbed symbol remains appropriately elliptic. Perturbations need to be smooth in order to apply the PDO techniques used in Secs. II and III. It is possible to use the stability of indices of Fredholm operators against compact perturbations and consider less regular perturbations as well, although we will not do so here.
For instance, we may consider with and bj(x) smooth, bounded below and above by positive constants, and say equal to 1 outside of a compact set in . Then we verify that the corresponding symbol though not necessarily in . Note that a more isotropic perturbation of the form for bj(X) smooth and equal to 1 outside of a compact set in would generate a perturbation in although one that is no longer a differential operator. This illustrates the reason why we considered the (reasonably large) classes .
The model Hamiltonian in the presence of one domain wall is h1(X) = (ξ1, …, ξd, x1). Based on Theorem 5.1, its topological charge is again given by deg h1 = 1. Domain walls of the form bj(x)x1 even with bj(x) = 1 outside of a compact domain no longer necessarily generate perturbations such that a1 remains in and are, therefore, not allowed in the theoretical framework of this paper. We may, however, replace x1 by m(x1) equal to x1 outside of a compact set. Inside that compact set, the level set m(x1) = 0 is then arbitrary.
For a time-dependent picture of how wavepackets propagate along curved interfaces for two-dimensional Dirac equations, see also Refs. 5, 7, and 8.
For concreteness and illustration, we spell out some details of the calculations of the indices in Theorems 2.3 and 5.1 when d = 1. We then have a0 = ξ = h0 while a = ξ − ix and a1 = ξσ1 + xσ2. We then observe that . In polar coordinates ξ = r cos θ and x = r sin θ, we observe that a−1da = r−1dr − idθ whose integral along the curve r = 1, gives −2πi, and hence Index Op a = 1 as a direct application of the Fedosov–Hörmander formula (10). We now observe that the index may be computed as in Lemma 5.2 from a1 = ξσ1 + xσ2 with and a* = ξ + ix so that a−*da* = r−1dr + idθ. This shows that whose appropriate integral gives the topological charge. Now, for h1 = (ξ, x). Therefore, as in Lemma 5.4, . We recognize in the integral of the latter form over the circle an expression for the degree of h1 written as the degree of the Gauss map, which to associates h1(X)/|h1(X)|. Using the expression (26) of the degree over the unit disk C gives deg h1 = 1 since ∇h1 = I2 at the unique point X = 0 where h1 = 0.
The regularized “Dirac” operator considered at the beginning of this section (this is not quite a Dirac operator as is not a second-order Laplace operator when η ≠ 0) is given by H1 = Op h1 · Γ1 with vector field so that . We observe that h1 = 0 only at the point (0, 0, 0), where the Jacobian is upper-triangular with diagonal entries equal to 1. Applying (26) and Theorem 5.1, we thus obtain that the topological charge of H1 equals 1 independent of ɛ and η as advertised.
The above orientation of the vector fields hk is natural in the context of topological insulators or superconductors, which are typically first written for spatially-independent coefficients. A different orientation helps to better display the invariance of the indices of Dirac operators across spatial dimensions (see Ref. 40, Proposition 19.2.9 for a related construction). We start with and then define iteratively
The above construction is an example of the more general structure,
where we verify that Index f♯g = Index f Index g. We apply it with g = Fn−1 and f = Dn − ixn. It is then straightforward to obtain that Index Fn = 1 for all n ≥ 1. We then observe that with spinor (1, 0, …, 0)t of dimension 2n−1.
B. Dirac operator with magnetic field
We now incorporate constant magnetic fields at infinity for magnetic potentials written in an appropriate gauge. Let us consider the case d = 2 for concreteness and the (minimal coupling) operator,
with A = (A1, A2) the magnetic vector potential and V a bounded scalar potential with compact support, say. The magnetic field is given by B = ∇ × A = ∂1A2 − ∂2A1. We choose the Landau gauge such that and for an arbitrary (smooth) compactly supported perturbation. In that gauge, we obtain that
is an operator H1 = Op a1 with for n1 = 2. Note that for H0 = Op a0, we do not have that a0 belongs to because of the presence of the unbounded magnetic potential. We would also not have that a1 belongs to if were chosen in the symmetric gauge, for instance. While physical phenomena have to be independent of the choice of a gauge, the appropriate functional setting to handle constant magnetic fields, and hence unbounded magnetic potentials, is not. With the above construction, we obtain that 2πσI(H1) = Index F = 1 for F = H1 − ix2, since the topological charge is given by
We could more generally consider a magnetic field with constant and opposite values as x1 → ±∞, for instance, with . The topological charge of H1 remains equal to 1. The magnetic field, therefore, has no influence on the topological charge in this setting.
Such a result should be contrasted with the very different outcome we obtain in Ref. 50 for the same model of magnetic potential but with a bounded domain wall μ(x1) [with x1σ3 replaced by μ(x1)σ3 in the above definition of H1]. In such a setting, both magnetic and mass confinements compete to generate asymmetric transport. Only when the bounded domain wall μ(x1) converges to sufficiently large values μ± as x1 → ±∞ (for a fixed magnetic field) do we retrieve that the asymmetric transport of 2πσI(H1) equals 1 (see Ref. 50, Theorem 2.1). Whereas replacing a bounded domain wall with an unbounded one is practically irrelevant when mass terms are the only confining mechanism, this is no longer the case when several confining mechanisms are present in the system.
C. Higher-order topological insulator
Let us consider the Weyl operator D · σ in dimension d = 3. As we mentioned in the introduction, the operator H2 = σ1 ⊗ D · σ + σ2 ⊗ Ix1 + σ3 ⊗ Ix2 generates a hinge in the third direction along which asymmetric transport is possible. With our choice of orientation, we have 2πσI(H2) = −deg(ξ1, ξ2, ξ3, x1, x2) = −1.
By implementing more general domain walls, an arbitrary number of asymmetric modes may be obtained. This is done by considering for ,
We thus deduce from Theorem 5.1 that
The last result is most easily obtained by identifying, as we did in the Proof of Lemma 5.5, the degree of h2 on the unit ball with the degree of the Gauss map from the unit circle to itself and then to the degree (winding number) of the map from the unit circle to itself, which equals p.
By an appropriate construction of the coefficients in the Hamiltonian H2 in (34) acting on , we thus obtain a low-energy model for a coaxial cable with an arbitrary number of asymmetric protected modes along the hinge (see, e.g., Ref. 52 for additional details on higher-order topological insulators).
D. Topological superconductors
Several superconductors and superfluids11,56 are modeled by Hamiltonians of the form
with coupling term HΔ = ∑i,j=1,2Δij(X)σi ⊗ σj for scalar operators Δij and for a mass of the quasi-particle m* > 0. For the above choice of the order parameter59 Δ, these Hamiltonians acting on separate into two 2 × 2 Hamiltonians (acting on the first and fourth components, and the second and third components, respectively). We now consider several such examples in one and two space dimensions.
E. One dimensional examples
For d = 1, an example with the order parameter Δ proportional to Dx gives
with . Let Δ = |Δ|eiθ and . We then verify that
and so we may assume Δ real-valued. Define and . We verify that , so that
This is of the form σ− ⊗ F* + σ+ ⊗ F with .
In order for F to be a Fredholm operator, we need to introduce a domain wall. This may be achieved in two different ways: it may be implemented by either the chemical potential μ = μ(x) or by the order parameter Δ = Δ(x). As we mentioned in the introduction of this section, the symbol of H1 has to be asymptotically homogeneous for the ellipticity condition (5) to hold and the theories developed in the preceding sections to apply. We thus regularize the operator using in (33). The regularization does not modify the symbol on compact domains in for 0 < ɛ sufficiently small and hence does not affect the computations of the index in (10) and (29).
When η > 0, we consider two regularized operators, one with a domain wall in the chemical potential,
and one with a domain wall in the order parameter
We observe that for H1 = Op a1, then is elliptic with m = 1 in the first example and m = 2 in the second example. Consider the second case (37). We wish to show that |h1|2 ≥ C(|X| − 1)4. This is clear for |x| ≤ 1 and for |x| ≥ 1, then |x| ≥ C⟨x⟩ɛ for C > 0, so that for C > 0. The latter expression is homogeneous in (ξ, ⟨x⟩ɛ) and non-vanishing on the unit sphere in these variables. This shows that . A similar computation shows that in (36). Note that we could also have used the following regularization for the first example: h1 = (ηξ2 − μx⟨x⟩ɛ, Δξ⟨ξ⟩ɛ), in which case .
We now compute the topological charges of the regularized operators starting with (36). We observe that h1(ξ, x) vanishes only at x = ξ = 0. The Jacobian there has a determinant equal to μΔ. The topological charge of (36) is, therefore, equal to Index F = deg h1 = sign(μΔ) assuming μΔ ≠ 0. Here and below, F is defined as usual by the relation H1 = σ− ⊗ F* + σ+ ⊗ F.
We next turn to (37), where h1(ξ, x) vanishes when ηξ2 = μ and x = 0. When μ < 0, there is no real solution to this equation, and the topological charge vanishes. When μ > 0, we have two solutions . At these points, the Jacobian matrix ∇h1 has components (2ηξ; 0; Δx; Δξ) with a determinant equal to 2ηΔξ2. The topological charge of H1 in (37) is, therefore, equal to Index F = deg h1 = 2 sign(Δ).
Let us finally consider the asymptotic regime η = 0 for a mass term m* → ∞ and a corresponding Hamiltonian H1 = −μσ1 + ΔDxσ2. A domain wall in the chemical potential is then modeled by μ(x) = μx. We then observe that H1 = Op a1 with and a topological charge equal to Index F = deg h1 = sign(μΔ) as in the setting η > 0. A domain wall in the order parameter requires the following regularized Hamiltonian , which is, however, gapped for μ ≠ 0 and hence topologically trivial.
F. Two-dimensional examples
We now consider two-dimensional examples of the above superconductor models. The p + ip (or p-wave) model with an order parameter proportional to momentum, is of the form
We assume here that Δ1 and Δ2 are real-valued. The case η = 0 is a Dirac operator and was treated earlier. We thus assume η > 0. A domain wall in the chemical potential is then implemented as
The symbol of this operator is h1 · Γ1 with . This regularization ensures that is elliptic. We could have defined a regularization in instead with .
It remains to compute the degree of h1. We find that h1 = 0 when ξ1 = ξ2 = x1 = 0 and that the Jacobian determinant there is equal to −μΔ1Δ2. The topological charge of the above operator is therefore 2πσI(H1) = deg h1 = − sign(μΔ1Δ2), which is consistent with Ref. 11.
In Ref. 56, Chap. 22, the domain wall is implemented in the order parameter Δ1, which, after appropriate regularization, gives
The constants Δ1, Δ2, and μ are assumed not to vanish and μ > 0. The symbol is given by
We have h1 = 0 when ξ2 = 0, x1 = 0, and . At each of the two solutions, the Jacobian of h1 is given by
Therefore, the topological charge of H1 is equal to 2πσI(H1) = deg h1 = −2 sign(Δ1Δ2) as in Ref. 56, Chap. 22. When μ < 0, the find deg h1 = 0 again.
In Ref. 56, Chap. 22, a model for a d-wave superconductor is given as
Following Ref. 56, Chap. 22, we implement a domain wall in Δ1(x1) and a regularization that gives the operator
This generates a symbol as may be verified. Then h1 = 0 when while x1 = 0. At each of these four roots, we compute
The sign of the Jacobian is the same at each of the roots so that deg h1 = 4 sign(Δ1Δ2). Therefore, we obtain a topological charge 2πσI(H1) = deg h1 = 4 sign(Δ1Δ2) when both μ > 0 and η > 0. When μ < 0, the operator is gapped and topologically trivial again.
G. Three-dimensional example
Following Ref. 11, (17.24), we consider the time-reversal invariant superconductor (or superfluid) model
acting on . When η = 0 and μ(X) = μx1, we obtain a standard Dirac operator with a topological charge sign(μΔ), as may be verified (see also the following calculations). When η > 0, we conjugate the above operator by g1 ⊗ I (which maps σ3 to −σ2), and after regularization and domain wall μ(X) = μx1, we obtain the Hamiltonian
The operator has an elliptic symbol in . We can then introduce as earlier H2 = H1 + σ3 ⊗ Ix2 and F = H2 − ix3. Following Theorem 5.1, the topological charge of H1 is then defined as Index F = 2πσI(H2) = −deg h1 with
We find h1 = 0 at the point ξ = 0 and x1 = 0. The Jacobian ∇h1 at this point has a determinant Δ3μ so that the topological charge is given by Index F = − sign(μΔ).
H. Other Hamiltonians
The above examples all fit within the framework of operators with symbols ak = hk · Γk verifying that is a scalar operator resulting in two energy bands. The ellipticity requirement is that the energies tend to infinity as |X| goes to infinity with a prescribed power m > 0. In this setting, the topological charge can conveniently be computed as the degree of the field hk as shown in the preceding examples.
The computations easily extend to operators of the form or more general direct sums of operators that are in the above form. More generally, the topological charge conservation result in Theorem 4.1 applies to operators beyond those of the form ak = hk · Γk provided that the symbol has eigenvalues appropriately converging to ∞ as |X| → ∞. For instance, the topological charge conservation in Theorem 4.1 applies to the sequence of effective Hamiltonians one obtains for continuous models of two-dimensional Floquet topological insulators. Such effective Hamiltonians are not in the form (23) and their asymmetric transport properties are most easily estimated by bulk-difference invariants related to the Fedosov–Hörmander formula (10) (see Ref. 9).
The theory presented in this paper does not apply for operators with flat bands for which the ellipticity condition cannot hold. A typical example is based on the shallow water wave (two-dimensional) Hamiltonian24,54
where f = f(x1) represents a (real-valued) Coriolis force. The symbol of that operator has two eigenvalues ±λ(x, ξ) with similar to those of a Dirac operator and a third uniformly vanishing eigenvalue. Therefore, H0 + α is gapped for α ≠ 0, but with a gap independent of ξ. The presence of this flat band in the essential spectrum creates difficulties that are not only technical; the topological charge conservation (a bulk-interface correspondence in dimension d = 2) in Theorem 4.1 does not always hold, although it does for certain profiles f(x) (see Refs. 6, 37, and 55).
A properly regularized version of the above Hamiltonian, however, does fit into the framework of the current paper. We observe that the kernel of a0 is associated with the eigenvector . We define the projector and for the regularized (pseudo-differential) Hamiltonian . The symbol now has eigenvalues given by ±λ and (ensuring that the symbol of Hμ is smooth). If we choose f(x1) = νx1 with ν ≠ 0 to generate a domain wall in the first variable, we observe that Hμ is elliptic with a symbol in (with m = 1). Following computations in, e.g., Refs. 6 and 49, which we do not reproduce here, we find that the topological charge of Hμ equals 2 sign(ν) independently of the choice of μ ≠ 0. This is the topological charge obtained when μ = 0 under smallness constraints in Ref. 6.
As we mentioned earlier, Hamiltonians with flat Landau levels do not satisfy the required ellipticity conditions. One such example is the ubiquitous two-dimensional scalar magnetic Schrödinger operator,
for instance, for A = (0, Bx1), so that ∇ × A = B is a constant magnetic field. For the numerous applications of this model, both discrete and continuous, to the understanding of the integer quantum Hall effect, we refer the reader to, e.g., Refs. 1, 2, 10, 22, 26, 29, 39, and 53. The spectral decomposition of this operator gives rise to a countable number of infinitely degenerate flat bands, the Landau levels, which are incompatible with the elliptic structure we impose on the symbol of the Hamiltonian in this paper. See Refs. 26 and 50 for the analysis of the transport asymmetry for Schrödinger and Dirac operators in the presence of magnetic domain walls. Note that the integral in (10) vanishes for a scalar-valued in dimension d ≥ 2 since then . Therefore, (10) would not predict the asymmetric transport observed in Refs. 26 and 50 in the presence of magnetic domain walls. Beyond ellipticity constraints, we finally remark that the edge conductivity (11) for magnetic Schrödinger (40) or Dirac operators is not invariant under semiclassical scaling D → hD.
ACKNOWLEDGMENTS
This work was funded in part by the NSF under Grant Nos. DMS-1908736 and EFMA-1641100.
AUTHOR DECLARATIONS
Conflict of Interest
The author has no conflicts to disclose.
Author Contributions
Guillaume Bal: Writing – original draft (lead).
DATA AVAILABILITY
Data sharing is not applicable to this article as no new data were created or analyzed in this study.
APPENDIX: NOTATION, OPERATORS AND FUNCTIONAL CALCULUS
This appendix recalls the results summarized in Ref. 15 allowing us to characterize spaces of symbols ak for Hk = Op ak adapted to operators modeling unbounded domain walls, domains of definition for Hk, as well as functional calculus showing, in particular, that and φ′(Hd−1) are pseudo-differential operators. We also recall results on semiclassical calculus and the Helffer–Sjöstrand formula following.25
On phase space in d spatial dimensions parameterized by X = (x, ξ) with and , we define a Riemannian metric g in the Beals–Fefferman form by
We use the notation for |·| the Euclidean norm applied to a vector u. For u = (u1, u2), we use the notation . Associated to the above metric g, we define the Planck function h(X) and its inverse λ(X) by 1 ≤ h−1(X) = λ(X) = Φξ(X)Φx(X).
In this paper, we consider two metrics: gi and gs. The metric gi is defined by with then hi = ⟨x,ξ⟩−2. The metric gs is defined by and with Planck function hs = ⟨x⟩−1⟨ξ⟩−1.
For 0 ≤ k ≤ d, we decompose with and . We define the weights
Following classical calculations,15 the weights Mk are admissible for g ∈ {gi, gs} in the sense of Ref. 15, Definition 2.3, and satisfy that Mk ≤ Cλp for some C > 0 and p < ∞ when λ ∈ {λi, λs}. This implies that MkhN goes to 0 as X → ∞ for N sufficiently large when h ∈ {hi, hs}.
For g = gs, a symbol b ∈ S(M, gs) when for each multi-index (α, β), we have
This is (4) when . For g = gi, a symbol b ∈ S(M, gi) when instead
The metric gi is referred to as the isotropic metric. We have that S(M, gi) a subspace of S(M, gs). Since they appear repeatedly in the derivations, we define for 0 ≤ k ≤ d the spaces,
We also define and . Here, nk and nd−1 are the dimensions of the spinors given in the introduction and in the construction of the augmented Hamiltonians in Sec. II, while is the space of n × n matrices with complex coefficients.
Associated to the weights Mk are the Hilbert spaces defined such that whenever a ∈ S(Mk, g). These spaces are independent of g ∈ {gs, gi}15 and hence referred to as . We observe that . The Hilbert spaces associated with and are denoted for 0 ≤ k ≤ d by
When , we thus obtain that Op ak maps to . Note that . Pseudo-differential operators with Weyl quantization are defined in (2) with integrals defined as oscillatory integrals.
Ellipticity (Ref. 15 , Sec. 2.3.3). We say that Hermitian valued is elliptic when
for some positive constants C1,2. This is equivalent to imposing that each eigenvalue of a(X) is bounded away from 0 by at least CM(X) outside of a compact set. We then say that and define the corresponding spaces of Hermitian elliptic symbols as for 0 ≤ k ≤ d and .
Since M(X) ≤ Cλp(X), ellipticity implies that is a self-adjoint operator with domain of definition and such that for some positive constant C,15
For , we thus obtain that Hk = Op ak is a self-adjoint operator from its domain of definition to . Similarly, for , then F = Op a is an unbounded operator from its domain of definition to .
Functional calculus (Ref. 15 , Sec. 2.3). For H = Op a and a elliptic, the above results show that the resolvent (z − H)−1 is an isomorphism from to for when Im(z) ≠ 0.
With the above assumptions, we have the Wiener property15 stating that: (i) A ∈ Op S(1, g) invertible in implies that A−1 ∈ Op S(1, g); and (ii) A ∈ Op S(M, g) bijection from to , then A−1 ∈ Op S(M−1, g).
This allows us to state the following result:
Let . Then is an isomorphism from to .
The above shows that maps to and has a symbol in . Moreover, using the Helffer–Sjöstrand formula as done in Ref. 15, Theorem 4 using p → −∞ in the notation there, we obtain the following result on the functional calculus:
Let and . Then .
The above result means the following in terms of seminorms: for each and each seminorm k defining the topology on the space of symbols, there is a seminorm l such that ϕ(H) is bounded for the seminorm k uniformly in the seminorm l applied to a. For a sequence of operators H(ɛ) = Op a(ɛ) with a(ɛ) with seminorms of uniformly bounded in ɛ, this implies that the symbol of ϕ(Hɛ) is bounded in any uniformly in ɛ as well.
Semiclassical calculus.25 The computation of several topological invariants, as in the proof of Ref. 40, Theorem 19.3.1, simplifies in the semiclassical regime. Let 0 < h ≤ h0 ≤ 1 be the semiclassical parameter. We define semiclassical operators in the Weyl quantization as
for a(X; h) a matrix-valued symbol in for each and h ∈ (0, h0] and ψ(x) a spinor in . The semi-classical symbol a(X; h) is related to the Schwartz kernel K(x, y; h) of Hh by
Note that Op a(x, hξ; h) = Opha(x, ξ; h). We define the classes of semi-classical symbols as Sj(M) constructed with the semi-classical metric in Beals–Fefferman form with Φx(X) = 1 and Φξ(X) = h−1, and for M an order function, i.e., in this context, a non-negative function on satisfying M(x, ξ) ≤ C(1 + |x − y| + |ξ − ζ|)NM(y, ζ) uniformly in (x, y, ξ, ζ) for some C(M) and N(M). Then if for each component b of a, we have for each 2d-dimensional multi-index α, a constant Cα such that
We will mostly use the case j = 0. We also use the notation b(X) ∈ S0(M) for symbols b(X) independent of h.
For two operators Opha and Ophb with symbols a ∈ S0(M1) and b ∈ S0(M2), we then define the composition Ophc = OphaOphb with symbol c ∈ S0(M1M2) given by the (Moyal) product (Ref. 25, Theorem 7.9),
For a ∈ S0(1), we obtain (Ref. 25, Theorem 7.11, Ref. 13, Proposition 1.4) that Oph a is bounded as an operator in with bound uniform in 0 < h ≤ h0 so that I − hOph a is invertible on that space when h is sufficiently small.
An operator is said to be semiclassically elliptic when the symbol a = a(x, ξ; h) ∈ S0(M) is invertible in for all and h ∈ (0, h0] with then a−1 ∈ S0(M−1).
Following Ref. 25 (see Ref. 6, Lemma 4.14), we obtain the following results on resolvent operators. Let Hh = Oph a with a ∈ S0(M). Let with ω ≠ 0. Then is a bounded operator and there exists an analytic function z → rz = rz(y, ζ; h) such that (compare to Lemma A.1). Moreover, the symbol rz ∈ S0(1) satisfies
for all multi-indices β = (βy, βζ) and a constant Cβ independent of z ∈ Z a compact set in and 0 < h ≤ h0.
Trace-class criterion. We have the following trace-class criterion (Ref. 25, Chap. 9). Assume that and that |∂αa(X)| ≤ CαM(X) for all |α| ≤ 2d + 1. Then Op a is a trace-class operator and
In other words, all symbols in S0(M) with M integrable generate trace-class operators.
Helffer–Sjöstrand formula.23,25 Finally, we recall some results on spectral calculus and the Helffer–Sjöstrand formula following Refs. 23 and 25; see also Ref. 13 for the vectorial case. For any self-adjoint operator H from its domain to and any bounded continuous function ϕ on , then ϕ(H) is uniquely defined as a bounded operator on (Ref. 25, Chap. 4). Moreover, for ϕ compactly supported, we have the following representation:
where, for z = λ + iω, d2z ≔ dλdω, , and where is an almost analytic extension of ϕ. The extension may be chosen as compactly supported in . Moreover, and on the real axis. We can choose the almost analytic extension such that for any uniformly in (λ, ω). Several explicit expressions, which we do not need here, for such extensions are available in Refs. 23 and 25.
REFERENCES
We use Δ for the order parameter as is customary in the superconductor literature. The (positive) Laplace operator is denoted by D · D.